Вы находитесь на странице: 1из 1844

Handbook of Advanced

Magnetic Materials
Volume I: Advanced Magnetic Materials:
Nanostructural Effects
Handbook of Advanced
Magnetic Materials
Volume I: Advanced Magnetic Materials:
Nanostructural Effects

Edited by:

Yi Liu
Center for Materials Research and Analysis
University of Nebraska
Lincoln, Nebraska

David J. Sellmyer
Center for Materials Research and Analysis
University of Nebraska
Lincoln, Nebraska

Daisuke Shindo
Institute of Multidisciplinary Research for Advanced Materials
Tohoku University
Sendai, Japan

(A) Tsinghua University Press ~ Springer


Library of Congress Cataloging-in-Publication Data

ISBN-IO: 1-4020-7983-4 e-ISBN-IO: 1-4020-7984-2


ISBN-13: 978-1402-07983-2 e-ISBN-13: 978-1402-07984-9

2006 Springer Science+Business Media, Inc.


All rights reserved. This work may not be translated or copied in whole or in part without the
written permission of the publisher (Springer Science+Business Media, Inc., 233 Spring Street, New
York, NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis.
Use in connection with any form of information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter developed is
forbidden.
The use in this publication of trade names, trademarks, service marks and similar terms, even if they
are not identified as such, is not to be taken as an expression of opinion as to whether or not they are
subject to proprietary rights.

Printed in the United States of America.

9 8 7 6 5 432 I SPIN 11097730

springeronline.com
Handbook of Advanced Magnetic Materials
Preface

In December 2002, the world's first commercial magnetic levitation supertrain


went into operation in Shanghai. The train is held just above the rails by
magnetic levitation (maglev) and can travel at a speed of 400 km/hr
completing the 30km journey from the city to the airport in minutes. Now
consumers are enjoying 50 GB hard drives compared to 0.5 GB hard drives ten
years ago. Achievements in magnetic materials research have made dreams of
a few decades ago reality. The objective of this book is to provide a
comprehensive review of recent progress in magnetic materials research. The
whole book consists of four volumes, each volume focusing on a specific field.
Graduate students and professional researchers are targeted as the readers.
Each chapter will have an introduction to give a clear definition of basic and
important concepts of the topic. The details of the topic are then elucidated
theoretically and experimentally. New ideas for further advancement are then
discussed. Sufficient references are also included for those who wish to read
the original work. Many of the authors are well known senior scientists. We
have also chosen some accomplished young scientists to provide reviews on
new and active topics.
In the last decade, one of the most significant thrust areas of materials
research has been nanostructured magnetic materials. There are several
critical sizes that control the behavior of a magnetic material. For example,
the coercivity of a magnetic material made of particles increases with
decreasing particle size, reaching a maximum where coherent rotation of a
single-domain particle is realized, and then decreases with further decrease of
the particle size. For a composite made of a magnetically hard phase and soft
phase, when the grain size of the soft phase is sufficiently large, the soft and
hard phases reverse independently. However, when the grain size of the soft
phase is reduced to a size of about twice the domain wall thickness of the hard
VI Preface

phase, the soft and hard phases will be exchange-coupled and behave as if a
single magnetic phase is present. Such behavior can be used to increase the
energy product of high-performance permanent magnets. Size effects become
critical when dimensions approach a few nanometers, where quantum
phenomena appear. The first volume of the book has therefore been devoted
to the recent development of nanostructured magnetic materials, emphasizing
size effects.
Our understanding of magnetism has advanced with the establishment of
the theory of atomic magnetic moments and itinerant magnetism. In general,
the magnetism of a bulk material can be considered as the superposition of
atomic magnetic moments plus itinerant magnetism due to conduction
electrons. In practical applications the situation becomes much more
complicated. The boundary conditions have to be taken into account. This
includes the size of the crystals, second-phase effects and intrinsic properties
of each phase. The effects of magnetic relaxation over long periods of time
can be critical to understanding. Simulation is a powerful tool for exploration
and explanation of properties of various magnetic materials. Simulation also
provides insight for further development of new materials. Naturally, before
any simulation can be started, a model must be constructed. This requires that
the material be well characterized. Therefore the second volume of the book
provides a comprehensive review of both experimental methods and simulation
techniques for the characterization of magnetic materials. After an introduction, each
section gives a detailed description of the method and the following sections
provide examples and results of the method. Finally further development of the
method will be discussed.
The success of each type of magnetic material depends on its properties
and cost which are directly related to its fabrication process. Processing of a
material can be critical for development of artificial materials such as
multilayer films, clusters, etc. Moreover, cost-effective processing usually
determines whether a material can be commercialized. In recent years
processing of materials has continuously evolved from improvement of
traditional methods to more sophisticated and novel methods. The objective of
the third volume of the book is to provide a comprehensive review of recent
developments in processing of advanced magnetic materials. Each chapter will
have an introduction and a section to provide a detailed description of the
processing method. The following sections give detailed descriptions of the
processing, properties and applications of the relevant materials. Finally the
potential and limitation of the processing method will be discussed.
The properties of a magnetic material can be characterized by intrinsic
Preface VB

properties such as anisotropy, saturation magnetization and extrinsic


properties such as coercivity. The properties of a magnetic material can be
affected by its chemical composition and processing route. With the continuous
search for new materials and invention of new processing routes, magnetic
properties of materials cover a wide spectrum of soft magnetic materials, hard
magnetic materials, recording materials, sensor materials and others. The
objective of the fourth volume of this book is to provide a comprehensive
review of recent development of various magnetic materials and their
applications. Each chapter will have an introduction of the materials and the
principals of their applications. The following sections give a detailed
description of the processing, properties and applications. Finally the
potential and limitation of the materials will be discussed.
NASA is considering the launching of spacecraft by maglev. The first
stage rocket, which accounts for two-thirds of the cost and is lost every
launch, would be replaced by a maglev track. Using a 50 ft track NASA
scientists have accelerated a model spacecraft to 96kph in less than half a
second. In the last few decades the knowledge of mankind has been expanding
rapidly into deep space measured by light years and the nano world where
building blocks of atoms are being engineered. Magnetism and magnetic
materials are among the most intriguing and fascinating science and
engineering fields. Undoubtedly advances in magnetic materials research will
continue to fuel our understanding of the universe in the new century. We hope
this book will provide a useful reference for researchers working at the frontier
of magnetic materials research.
We would like to express our sincere thanks to all our devoted authors,
technical editors, and publishers for making this book possible.

The editors
Contents

Preface V
List of Contributors XVI

1 Intrinsic and Extrinsic Properties of Advanced Magnetic Materials . .. . ... 1


1. 1 Introduction ... ... ... 1
1.2 Intrinsic Properties 3
1. 2. 1 Magnetic Moment 3
1. 2. 2 Exchange'" ... ... ... 4
1. 2. 3 Magnetization and Magnetic Order 6
1. 2. 4 Itinerant Magnetism 8
1. 2.5 Magnetic Anisotropy 10
1. 3 Extrinsic Properties ... ... 14
1. 3. 1 Coherent Rotation 16
1.3.2 Domains and Domain Walls 18
1.3.3 Coercivity 21
1. 4 Magnetic Materials 25
1. 4. 1 Permanent Magnets 25
1. 4. 2 Soft Magnets 27
1. 4.3 Recording Media 28
1. 4.4 Other Magnetic Materials 29
1.4.5 Tables 29
1. 5 Magnetic Nanostructures 31
1. 5. 1 Physical Classification of Magnetic Nanostructures 32
1.5.2 Intrinsic Properties and Finite-Size Effects 34
1. 5.3 Narrow-Wall and Constricted-Wall Phenomena 36
1. 5. 4 Nanomagnetic Localization 37
1. 5. 5 Cooperative Effects ... ... ... 40
1. 5. 6 Random Anisotropy and Remanence Enhancement 42
1. 5. 7 Magnetization Dynamics in Nanostructures 43
1. 5. 8 Energy-Barrier Laws 47
1.6 Conclusions .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. 49
Appendix 1. 1 Magnetic Units 50
References 50

2 Magnetism in Ultrathin Films and Beyond 58


2. 1 Introduction .. 58
2.2 Fabrication 59
X Contents

2.2. 1 Ultrathin Films 59


2.2.2 Wires 61
2. 2. 3 Dots 62
2.3 Magnetic Properties in Low Dimensional Systems 64
2.3. 1 Metastable Structures 64
2.3.2 Dimensionality and Phase Transition 66
2.3.3 Surface/Interface Electronic Structure 67
2.3.4 Quantum Size Effects 69
2.3.5 Domains and Domain Walls 71
2. 4 Conclusions'" 74
References 74

3 Classical and Quantum Magnetization Reversal Studied in


Nanometer-Sized Particles and Clusters 77
3. 1 Introduction .. .. .... .... .. .. .. .. .. .. .. .. 77
3.2 Single-Particle Measurement Techniques 79
3. 2. 1 Overview of Single-Particle Measurement
Techniques 79
3.2.2 Micro-SQUID Magnetometry 81
3. 3 Mechanisms of Magnetization Reversal at Zero Kelvin 82
3. 3. 1 Magnetization Reversal by Uniform Rotation
(Stoner-Wohlfarth Model) 82
3.3.2 Nonuniform Magnetization Reversal...... 97
3. 4 Influence of Temperature on the Magnetization Reversal 102
3. 4. 1 Neel-Brown Model of Thermally Activated Magnetization
Reversal ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 102
3.4.2 Experimental Methods for the Study of the Neel-Brown
Model 103
3. 4.3 Experimental Evidence for the Neel-Brown Model 106
3. 5 Magnetization Reversal by Quantum Tunnel ing 112
3. 5. 1 Quantum Tunnel ing of Magnetization in Molecular
Clusters ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 113
3.5.2 Quantum Tunneling of Magnetization in Individual Single-
Domain Nanoparticles 116
3.5.3 Magnetization Measurements of Individual Single-Domain
Nanoparticles and Wires at Very Low
Temperatures 118
3. 5.4 Quantization of the Magnetization 120
3.6 Summaries and Conclusions 122
References 123

4 Micromagnetic Simulation of Dynamic and Thermal Effects 128


4. 1 Introduction ... ... ... ... ... ... 128
4.2 Micromagnetic Background 129
Contents XI

4. 2. 1 Equation of Motion 129


4. 2. 2 Gibbs Free Energy 130
4.2. 3 Langevin Equation 131
4.2. 4 Characteristic Length Scales 132
4. 3 Numerical Techniques ... ... ... ... ... ... 133
4.3. 1 Finite Element Discretization 133
4.3.2 Magnetostatic Field Calculation 134
4. 3. 3 Time Integration 135
4.4 Numerical Examples 137
4. 4. 1 Small Particles 137
4.4.2 Thin Film Elements 140
4. 4. 3 Circular Nanodots 140
4.4.4 Magnetic Nanowires 142
References 145

5 Magnetic Relaxation and Quantum Tunneling of Magnetization 147


5. 1 Introduction 147
5. 2 Magnetic Relaxation and Related Phenomena in Monosized,
Non-Interacting Particle Systems .. .. .. .. 148
5.2. 1 Introduction ... ... ... ... 148
5.2.2 Blocking Temperature'" 149
5.2.3 a.c. Magnetic Susceptibility Measurement 150
5. 2. 4 Zero-F ield-Cooled and Field Cooled Magnetization
Curves 155
5. 2. 5 Magnetic Hysteresis Loops 158
5.2.6 Numerical Simulation Results 159
5. 3 Magnetic Relaxation in Particle Systems with Size Distribution 166
5. 3. 1 Introduction ... ... ... ... ... ... ... ... 166
5.3.2 Logarithmic Magnetic Relaxation 167
5.4 Quantum Tunneling of Magnetization 171
5. 4. 1 Introduction ... ... ... ... 171
5. 4.2 Physics Related to QTM 172
5. 4 . 3 Observation of QTM in Systems of Nanostructured
Materials 175
5. 5 Conclusions and Future Perspectives 177
References ... ... ... ... ... ... ... ... ... ... ... ... ... 177

6 Nanostructured Exchange-Coupled Magnets 182


6. 1 Introduction ... ... ... ... ... ... ... ... ... ... 182
6.2 Theory of Exchange-Coupled Magnets 184
6.2. 1 Energy Product ... ... ... 184
6.2.2 Fundamental Equations 186
6.2.3 Nucleation Field 188
6. 2. 4 Energy Product 189
XII Contents

6.2.5 Micromagnetic Localization 190


6.2.6 Texture'" ... ... ... ... 191
6.2. 7 Effective Exchange 192
6. 3 Experimental Systems ... ... 196
6.3. 1 FePt-Based Magnets 196
6. 3.2 Rare-Earth Cobalt Magnets 207
6. 3. 3 Nd-Fe-B-Based Magnets 220
6.4 Conclusions 254
References ... ... ... 255

7 High-Field Investigations on Exchange Coupling in R-Fe Intermetallics


and Hard/Soft Nanocomposite Magnets 267
7. 1 Introduction 267
7.2 Exchange and Crystal Field Model for (R, R' )-Fe-X
System 269
7.3 High-Field Magnetization, Spin Reorientation and
Magnetostriction in (R, R')2 Fe14B 271
7.3. 1 Magnetic Phase Diagram and Spin Reorientation in
(Erl-x Tbx)2Fe14B 271
7.3.2 Magnetostriction and Spin Reorientation in
(Er,-x Tbx)2Fe,4B 274
7.4 Exchange-Coupling in Hard/Soft Nanocomposite Films 278
7.4. 1 Multilayerd Nd2Fe14 B/ex-Fe Films 278
7.4.2 Nanodispersed Nd2Fe14 B/ex-Fe Films 284
7.5 Conclusions 290
References ... ... ... 291

8 Fabrication and Magnetic Properties of Nanometer-Scale Particle


Arrays 294
8. 1 Introduction 0 294
8.2 Fabrication of Regularly Arranged and Shaped Particles
on a Nanometer Scale 295
8. 2. 1 Overview'" ... 295
8.2.2 Scanning Tunneling Microscope Assisted Chemical
Vapor Deposition 296
8. 3 Magnetic Measurements on Extremely Small Particles 299
8. 3. 1 Overview .. ... ... 299
8. 3.2 Hall Gradiometry 301
8. 3. 3 Optimized Hall Measurements for Particle Arrays 304
8.3.4 Variable Field Magnetic Force Microscopy (MFM) in
Nanoparticles 307
8. 4 Magnetization Processes in Small Particles 311
8. 4. 1 Reversible Rotation of Magnetization 311
Contents XIII

8.4.2 Models for Magnetization Reversal without Thermal


Activation 316
8.4.3 Experimental Results on Magnetization Reversal 316
8.4.4 Interaction Effects ... ... ... ... ... ... ... ... 317
8.4.5 Comparison to Numerical Simulations 320
8 . 5 Thermally Activated Magnetization Reversal in Small Particles 322
8. 5. 1 General Considerations 322
8. 5.2 Phenomenological Model 323
8. 5. 3 Angular Dependence of Magnetization Reversal 327
8. 5. 4 Magnetic Viscosity 328
8.6 Conclusions 331
References '" 331

9 Processing and Modeling of Novel Nanocrystalline Soft Magnetic


Materials 339
9. 1 Introduction 339
9.2 Origin of Magnetic Softness-Random Magnetocrystalline
Anisotropy 341
9.2. 1 Magnetic Anisotropies and Magnetic Softness 341
9.2.2 Random Magnetocrystall ine Anisotropy 345
9.3 Nanostructure-Magnetic Properties Relationships 351
9. 3. 1 Grain Size and Magnetic Softness .... .. .. 351
9.3.2 Intergranular Phase and Magnetic Coupling 357
9.3.3 Application-Oriented Magnetic Properties 360
9. 4 Principles Underlying Alloy Design 364
9.4. 1 Alloying Elements and Alloy Systems 364
9.4.2 Alloy Design in Fe-Metal Based Nanocrystalline
Alloys 367
9.4.3Alloy Design in Fe-Metalloid Based Nanocrystalline
Alloys 368
9. 5 Prospects 370
References ... ... 371

Index 374
List of Contributors

1. R. Skomski Dept Phys & Astronomy and CMRA, UNL


D. J. Sellmyer Lincoln, NE 68588
rskomski@unlserve. unl. edu
E-mail: dsellmye@unlnotes. unl. edu
2. Dr. Dongqi Li Materials Science Division
Argonne National Lab
Argonne, IL 60439, USA
E-mail: dongqi@anl. gov
3. Wolfgang Wernsdorfer Lab. L. Neel - CNRS, BP166,
38042 Grenoble Cedex 9, France,
E-mail: wernsdor@labs.polycnrs-gre.fr
4. T. Schefl Institute of Applied and Technical Physics
J. Fidler Vienna University of Technology
D. Sness Wiedner Haupstr. 8- 10
V. Tsiantos A-1040 Wien
E-mail: Schrefl@atp330.tuwien.ac.at
5. Prof. Xixiang Zhang Department of Physics and Institute of Nano Sci-
ence and Technology (INST)
The Hong Kong University of Science and Tech-
nology
Clear Water Bay, Kowloon, Hong Kong, CHINA
E-mail: phxxz@ust.hk
6. Wei Lin Shenyang National Laboratory for Materials Sci-
ence and International Centre for Materials Phys-
ics, Institute of Metal Research, Chinese Acade-
my of Sciences, Shenyang 110016, P. R. China
E-mail: wliu@imr.ac.cn
Y. Lin Center for Materials Research and Analysis,
R. Skomski University of Nebraska, Lincoln, NE
D. J. Sellmyer 68588-0113, USA
XVI List of Contributors

7. Hiroaki Kato Department of Applied Physics, Graduate School


of Engineering, Tohoku University, Aoba-yama
08, Aoba-ku, Sendai 980-8579, Japan
Mitsuhiro Motokawa Institute for Materials Research, Tohoku Universi-
ty, 2-1-1 Katahira, Aoba-ku,
Sendai 980-8577, Japan
8. S. Wirth Max-Planck-Institute for Chemical Physics of Sol-
ids, N6thnitzer Str. 40, 01187
Dresden, Germany
S. von Molnar MARTECH , Florida State University, Tallahas-
see, FL 32306-4351, USA

9. Kiyonori Suzuki School of Physics and Materials Engineering,


Monash University, Victoria 3800, Australia
1 Intrinsic and Extrinsic Properties of Advanced
Magnetic Materials

R. Skomski and D. J. Sellmyer

1. 1 Introduction

Magnetic materials have inspired human imagination for millennia, and for
many centuries they have stimulated progress in science and technology. For a
long time, focus has been on naturally occurring magnetic materials, such as
iron and magnetite (see Fig. 1. 1). In the last few decades, there has been a
revolution in the development of magnetic materials. On one hand, atomic-
scale quantum-mechanical and relativistic effects have been exploited to
create high-performance magnetic materials, such as the alloys SmCos and
Nd2 Fe14 B, which are used to produce permanent magnets. On the other hand,
geometrically well-defined nanostructures such as multilayers, particle arrays
and bulk composites, are now actively explored and used to fabricate
magnetic materials for a wide range of appl ications (Coehoorn et aI., 1988;
Baibich et al. , 1988; Skomski and Coey, 1993; McCurrie, 1994; Himpsel et
al. , 1998; Comstock, 1999; Wood, 2000; Weller et aI., 2000; Ziese and
Thornton, 2001; Sellmyer et al. , 2002).
In magnetism, there is a fundamental distinction between intrinsic and extrinsic
properties. Intrinsic properties, such as the spontaneous magnetization M s ' the
Curie temperature T c' and magnetocrystalline anisotropy, are realized on atomic
length and time scales but describe infinite crystals. They can, in general, be
considered as equilibrium properties. For example, the magnetization of ex-Fe single
=
crystals, J-Io Ms 2. 15 T, is associated with the body-centered cubic structure of
elemental iron. (For magnetic units, see Appendix 1. 1.) By contrast, extrinsic
magnetic properties, such as the coercivity He and the remanence Mr , reflect the
magnet's real-structure (morphology) (Bloch, 1932; Landau and Lifshitz, 1935;
Kersten, 1943; Skomski and Coey, 1999). The strong real-structure dependence of
extrinsic properties is seen, for example, from the fact that the coercivity of
technical iron doubles by adding 0.01 wt. % nitrogen (Kersten, 1943). By
comparison, intrinsic properties are not affected by small concentrations of defects.
Extrinsic phenomena are, in general, nonequilibrium phenomena, closely related to
magnetic hysteresis.
2 R. Skomski and D. J. Sellmyer

Figure 1. 1 Popular magnetism. Compiled from 20th-century middle-school textbooks,


this figure correctly illustrates some phenomenological aspects of magnetism (magnetic
field, Zeeman interaction, flux closure, difference between hard and soft magnets) but
ignores the atomic origin of intrinsic magnetism (magnetic moment, magnetocrystalline
anisotropy) and the nanoscale origin of hysteresis (granular anisotropy fluctuations, grain
boundaries) .

The involvement of nanoscale phenomena leads to the question: How


many atoms are necessary to form a magnetic body, which can be considered
as quasi-infinite? The answer depends on the considered magnetic property.
Intrinsic properties tend to approach their bulk values on fairly small length
scales. For example, "long-range" thermodynamic fluctuations, as involved in
the realization of the Curie temperature, and deviations from the Bloch
character of metallic wave functions yield only small corrections when the size
of the magnetic particle exceeds about 1 nm. In the case of extrinsic
properties, nanostructural effects are important on much larger length scales,
typically at least several nm. One example is the enhancement of permanent
magnet energy-product in exchange-coupled hard-soft nanocomposites
(Skomski and Coey, 1993), for example in Fe-Pt composite films (Liu et al. ,
1998). This effect, where adding a soft phase improves the permanent-
magnet performance of a hard phase, is real ized on length scales of the order
of 10 nm.
The scope of this chapter is to provide an introduction into the wide range
of physical properties, nanogeometries, chemical compositions, and
applications of magnetic materials of current interest in science and
technology. Emphasis is on the physical nature of the phenomena. Numerical
calculations and computer experiments (Daalderop et al., 1990; Coehoorn,
1989; Bland and Heinrich, 1994; Schrefl et al., 1994; Steinbeck et al. ,
1996; Sabiryanov and Jaswal, 1998a, 1998b; Sandratskii, 1998; Schrefl and
Fidler, 1998; Lyberatos, 2000; Hertel, 2001) are of great importance in
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 3

advanced magnetism, but the discussion of numerical details goes beyond the
scope of this chapter. Section 1. 2 deals with the atomic aspects of
magnetism; Section 1. 3 is devoted to the phenomenon of hysteresis; Section
1. 4 briefly discusses the main classes of magnetic materials; and Section 1. 5
is devoted to nanomagnetic effects.

1. 2 Intrinsic Properties

Intrinsic properties refer to the atomic origin of magnetism and involve quantum
phenomena such as exchange, crystal-field interaction, interatomic hopping
and spin-orbit coupl ing (Ising, 1925; Heisenberg, 1928; Bloch, 1929;
Brooks, 1940; Slater, 1953). The understanding of some problems of intrinsic
magnetism, such as exchange (Heisenberg, 1928), dates back to the very
early days of quantum mechanics, whereas the understanding and exploitation
of the large magnetic anisotropy of advanced magnetic materials is a
comparatively recent event. Intrinsic properties themselves are interesting
figures of merit, but they also affect the hysteresis loop by entering the
micromagnetic equations as parameters. In this section, we focus on the
magnetic dipole moment per atom, the spontaneous magnetization, the
magnetocrystalline anisotropy and the exchange stiffness.

1. 2.1 Magnetic Moment

The moment of magnetic solids nearly exclusively originates from the partly
filled inner electron shells of transition-metal atoms. Of particular importance
are the iron-series transition-metal elements (3d elements) Fe, Co and Ni,
and the rare-earth or 4f elements, such as Nd, Sm, Gd and Dy. On the other
hand, palladium-series (4d), platinum-series (5d), and actinide (5f)
elements, have a magnetic moment in suitable crystalline environments. The
inner-shell electrons give rise to a magnetic moment m, which is often
measured in Bohr magnetons per formula unit (IJs = 9.2740 X 10- 24 Am 2 ). An
alternative way of characterizing a material's net moment is to consider the
spontaneous magnetization M s = mlV, measured in Aim, or its flux-density
equivalent IJo M s ' measured in T. Here V is a small volume element containing
at least one unit cell.
There are two sources of the atomic magnetic moment m: currents
associated with the orbital motion of the electrons (orbital moment J) and the
electron spin (spin moment s). The magnetism of free atoms or ions is
governed by Hund's rules, which predict the spin and orbital moment as a
function of the number inner-shell electrons (K ittel, 1986). Hund' s rules are
4 R.Skomski and D. J. Sellmyer

very well satisfied in rare-earth atoms, because the radius of the rare-earth 4f
shells (about O. 5 A or O. 05nm is much smaller than the atomic radius of the
rare-earth atoms (about 1. 8 A). The electrostatic field created by the
crystall ine envi ronment is therefore largely screened, and the rare-earth 4f
shells can be treated as quasi-free.
The magnetic moment of iron-series transition-metal atoms in metals (Fe,
Co, Ni, YCo s ) and nonmetals (Fe 3 04' NiO) is given by the spin, so that the
moment, measured in /-is' is equal to the number of unpaired spins. For
example, Fe 2 + (ferrous iron) has four unoccupied 3d ... orbitals, so that the
moment per ion is 4 /-is. The orbital moment is very small, because the orbital
motion of the electrons is suppressed or quenched by the electrostatic crystal
field. In terms of elementary quantum mechanics, 3d wave functions of free
atoms have a "circular-current" running-wave character and yield an orbital
moment, but a crystall ine environment forces the electrons to form standing
waves with zero orbital moment. On the other hand, spin-orbit coupling
(Section 1.2. 5) competes with the crystal field and yields a small admixture
of circular-current character, corresponding to a residual orbital moment of the
order of O. 1 /-is. Figure 1. 2 illustrates the charge density of unquenched and
quenched orb itals.

(a) (b)

Figure 1. 2 Charge distribution of electron clouds in inner shells of transition-metal atoms:


(a) unquenched, as in rare-earth ions such as Nd3+ , and (b) largely quenched, as in the
case of iron-series transition metals. Four-fold symmetries of the type (b) reflect, for
example, the crystal-field splitting between xy and x' - y'd orbitals in cubic and tetragonal
crystals. The degree of quenching depends on the ratio of crystal-field interaction and spin-
orbit coupling and determines the orbital moment and the magnetocrystalline anisotropy.

1. 2. 2 Exchange

The magnetic moment and the spontaneous magnetization are realized by the
exchange interaction between electrons. In a simple two-electron picture,
exchange gives rise to tt
(ferromagnetic) or H (antiferromagnetic)
coupling between spins. There are two main types of exchange. First, atomic
moments are determined by intra-atomic exchange. For example, the above-
mentioned ferric iron has six 3d electrons, and intra-atomic exchange yields
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 5

the schematic spin structure ttttH. Second, there is an interatomic


exchange between neighboring magnetic atoms. Interatomic exchange yields,
for example, the long-range magnetic order observed in ferromagnets
(Section 1. 2. 3), ensures finite-temperature magnetocrystalline anisotropy
(Section 1. 2. 5), and is of importance in micromagnetism (Section 1.3.2).
Exchange is an electrostatic many-body effect, caused by 1/1 r - r' I
Coulomb interactions between electrons located at rand r'. Physically, H
electron pairs in an atomic orbital are allowed by the Pauli principle but
unfavorable from the point of view of Coulomb repulsion. Parallel spin
alignment, tt, means that the two electrons are in different orbitals, which is
electrostatically favorable. However, the corresponding gain in Coulomb
energy competes against an increase in one-electron energies: only one
electron benefits from the low ground-state energy - the second electron must
occupy an excited one-electron level.
In agreement with Hund' s rules, intra-atomic exchange favors parallel
spin alignment. The sign of the interatomic exchange is more difficult to
predict. In the case of two electrons and two atomic sites, lowest-order
perturbation theory yields (Skomski and Coey, 1999)

(1. 1)

where U is the energy necessary to add a second electron into an atomic


orbital (Coulomb energy), T denotes the interatomic hopping integral, and J D
is the direct exchange (exchange integral). The direct exchange is always
positive, but for typical solid-state interatomic distances it is not larger than
about O. 1 eV, that is smaller than U by at least one order of magnitude. On
the other hand, hopping reduces the effective exchange, making Jeff less
ferromagnetic. In oxides, T U, and Eq. (1. 1) yields Jeff = J 0 - 2 T 2 / U. Due
to the smallness of the direct exchange, oxides are often antiferromagnets, but
when T = 0 by symmetry, as in Cr02' then J o gives rise to ferromagnetism
(Goodenough-Kanamori rules). In metals, the interatomic hopping is
proportional to the bandwidth, and ferromagnetism is real ized in narrow bands
(Section 1.2.4).
A widely-used approach to discuss interatomic exchange is the
Heisenberg interaction - Js 1 82 between neighboring spins 81 and 82' where
J is some effective exchange. One example is the Ruderman-Kittel or RKKY
interaction Ruderman-Kittel-Kasuya-Yosida interaction J -- cos (2k F R) / R 3
between local moments embedded in a free-electron gas of Fermi-wave vector
k F (Ashcroft and Mermin, 1976). Interatomic exchange competes not only with
finite-temperature disorder but also with micromagnetic magnetization
inhomogenities, such as domains. On a continuum level (Sections 1. 3. 2,
1. 3. 3, 1. 5. 2), the Heisenberg exchange translates into the energy density
Ad(V M)2/M~ where A is the exchange stiffness. For typical ferromagnets,
Ad is of the order of 10 pJ/m [1O- 11 J/m, 1O- 6 erg/cm (1 erg= 10- 7 J)]'
6 R.Skomski and D.J.Sellmyer

1. 2. 3 Magnetization and Magnetic Order

The term magnetic order refers to the atomic spin structure (ferromagnetic,
ferrimagnetic, antiferromagnetic, etc.); it usually excludes micromagnetic
structures, such as domains and domain walls. Interatomic exchange favors
parallel or anti parallel alignment of neighboring spins, respectively, depending
on whether the sign of the exchange constant J is positive or negative. In
ferromagnets, such as Fe, Co and Ndz Fe14 B, all spins are parallel and the
atomic moments add. Ferrimagnets, such as Fe 3 04 and BaFe 1Z 019' and
antiferromagnets, such as CoO and MnF z , are characterized by two (or more)
sublattices with opposite moments. This amounts to a ferrimagnetic reduction
or antiferromagnetic absence of a net moment. Sub lattice formation may be
spontaneous, as in typical antiferromagnets, or imposed by the atomic
composition, as in ferrimagnets. Competing exchange interactions in periodic
crystals and in disordered magnets give rise to noncollinear spin
arrangements. Examples are helimagnetic order in perfect crystals, which is
caused by competing interactions between next and more distant neighbors,
and spin-glass behavior in magnets with atomic-scale disorder (Moorjani and
Coey, 1984). Deviations from parallel or anti parallel spin alignment may also
occur at surfaces and interfaces. Figure 1. 3 shows some schematic examples
of magnetic order.

\ ! \ ! \ \ I I \ I ! \
\ \ ! ! I I \ I \ I \ I
I \ I \ I I \ I \ I \ I
I \ ! \ ! ! \ ! I ! \ I
I \ ! \ ! \ \ I I \ \ I
! I ! I \ ! I ! I \ I !
(a) (b)

\' I I I \ I I
- '. I , I
~

I
I
I
\
I
I
I
!
I
I
!
!
I
"' \
,-
--"';1
;I I
\ I ;I ~
I ! I \ I \ \-\ I I d5J
!
I
I
!
!
I
I
!
I
I
I
\
I '. \ I \
-I \ \
1j-!
~;I
"'
(e) (d)

Figure 1. 3 Magnetic order: Ca) ferromagnet at T = 0, Cb) antiferromagnet at T = 0,


Cc) ferrimagnet at T = 0, and Cd) ferromagnet above T e . At zero temperature,
noncollinear spin configurations such as Cd) are realized for example in spin glasses.

It is important to distinguish between zero-temperature magnetism and


finite-temperature magnetic order. The competition between interatomic
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 7

exchange and thermal disorder leads to the vanishing of the spontaneous


magnetization at a well-defined sharp Curie temperature T c. The total
interatomic exchange per atom does not exceed about O. 1 eV, corresponding
to Curie temperatures of the order of 1000 K. By contrast, from the ratio
/-Is/ k s = 0.672 K/T, it follows that typical magnetostatic fields in solids,
which are of the order of one tesla, cannot explain magnetic order at
temperatures above about 1 K.
Intra-atomic exchange is much larger than interatomic exchange, of the
order of 1 eV, so that atomic moments at Teare close to their zero-
temperature values. CD The relative strength of the intra-atomic exchange
means that typical magnetization changes in magnetic solids are caused by
moment rotations rather than by changes in the moments' magnitude., This
refers not only to the spontaneous magnetization but also to micromagnetic
magnetization changes. Figure 1. 3d illustrates this point by showing the spin
structure of a typical ferromagnet above the Curie temperature.
Due to thermal excitations, the spontaneous magnetization M s is
temperature-dependent. M s (T) is often but not always smaller than the zero-
temperature magnetization Ms (0) = Mo. Figure 1. 4 shows typical temperature
dependences M s ( T). By definition, M s (T) implies that the external magnetic
field is zero. Nonzero external fields smooth the temperature dependence of
the magnetization and lead to ill-defined Curie temperatures.
Ms
Spin waves

(a) (b) (c)

Ms

M'~
Tc T T

(d) (e) (f)

Figure 1. 4 Typical temperature dependences of the spontaneous magnetization:


Ca) simple ferromagnet, Cb) two-sub lattice ferromagnet, Cc) ferrimagnet without
compensation point, Cd) two-sublattice spin orientation, Ce) ferrimagnet with compensation
point and CO two-sublattice origin of the compensation point. Cd) explains the mechanism behind
Cb- c).

CD There are a few exceptions, such as very weak itinerant ferromagnets (for example
ZrZn2) and low-spin high-spin transition in fcc iron.
8 R.Skomski and D.J.Sellmyer

The derivation of the spontaneous magnetization M s of a solid from the


corresponding Heisenberg Hamiltonian is a complicated problem. Since the
spontaneous magnetization is an equilibrium quantity, the knowledge of the
partition function Z = ~[exp(- E"lk s nJ is sufficient to derive M s ' but the

number of terms in Z increases exponentially with the size of the magnet, and
only in a few cases there exist exact solutions (Yeomans, 1992).
The simplest finite-temperature approach is the spin-1/2 mean-field Ising
model. It is defined in terms of the two energy levels E =- hs , where
s = 1 is the orientation of the atomic spin and h = fJo mH is the exchange-
interaction field (mean field) acting on the atomic spin. It is an easy exercise
=
to find the thermally averaged spin projection <s > tanh( hi k s n. In mean-field
theory, h = zJ <s >, where z is the number of nearest neighbors and J is the
interatomic exchange, so that < s > = tanh (zJ < s >1k s T). The temperature
dependence of M s ( n = M o <s > determined from this self-consistent equation
is similar to Fig. 1. 4a, and the Curie temperature is equal to zJ 1k s The
Heisenberg model, which takes into account the vector character of the
quantum spins, has the mean-field Curie-temperature T c = (S + 1) zJ 13k s S ,
where S is the spin quantum number.
The mean-field model is easily generalized to two or more sublattices; for
N sublattices (or N non-equivalent atomic sites) it yields N coupled algebraic
equations (Smart, 1966; Skomski and Sellmyer, 2000). Figures 1. 4b - f
show some schematic examples. On the otherhand, as illustrated in Fig. 1. 4a,
the mean-field model does not work very well at low temperatures, where M s
is determined by cooperative spin waves (Bloch, 1930), and close to T c ,
where long-range critical fluctuations interfere (Ising, 1925; Brush, 1967;
Yeomans, 1992). This leads, for example, to the physically unreasonable
prediction of ferromagnetism in one dimension (Ising, 1925; Brush, 1967).

1. 2. 4 Itinerant Magnetism
Up until now, we have assumed that the magnetic moment is associated with
well-defined atomic sites. This local-moment picture is realized, for example,
in insulating transition-metal oxides and rare-earth metals. However, in the
iron-series transition-metal elements and in many transition-metal alloys, such
as MnBi, PtCo and ZrZn2, the magnetism is itinerant, that is, caused by
delocalized electrons. Itinerant magnetism is characterized by generally non-
integer moments. For example, the moments of Fe, Co and Ni are 2.2 fJs'
1. 7 fJs and O. 6 fJs per atom. The reason is the band-structure character of
itinerant magnetism: Due to interatomic hopping, the atomic 3d states were
broadened into bands, and the magnetic moment is realized by continuously
filling the t and ~ bands with electrons.
Nonmagnetic metals (Pauli paramagnets) have two equally populated t
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 9

and subbands. An applied magnetic field may transfer a few electrons from
the band to the t band, but straightforward calculation shows that the
corresponding spin polarization is very small, typically less than O. 1% . <Z>
Itinerant ferromagnetism is caused by atomic exchange fields, which are much
larger than laboratory-scale magnetic fields. Some guidance is provided by
Eq. C1. 1). In the metallic limit of strong interatomic hopping, the effective
exchange is roughly equal to U /4- T. This means that ferromagnetism occurs
for small interatomic hopping, that is, for narrow bands CSection 1. 2. 2). A
slightly more detailed analysis yields the Stoner criterion 1 DCE F l, where
1- U/ 4 is the Stoner parameter and 0 CE F) -1/ T is the paramagnetic density
of states (DOS) at the Fermi level CCoehoorn, 1989). For 3d elements, I~
1 eV, and 0 CE F) is sufficiently large to ensure ferromagnetism in Co, Ni and
bcc Fe. Some other elements, such as Pd and Pt, are very close to satisfying
the Stoner criterion and easily develop a magnetic moment in ferromagnetic
alloys. Figure 1. 5 shows typical spin-polarized DOS of itinerant ferromagnets.
The inset in the iron figure shows the paramagnetic DOS of bcc Fe: Since the
Fermi level lies in the main peak, the Stoner criterion is satisfied.

Fe (bee) Co (hep)
t

Ni

D(E)
LE(a.u.)

Figure 1. 5 Densities of states (DOS) for some ferromagnets and for fictitious Pauli-
paramagnetic bcc Fe (inset). The electrons fill the band up to the Fermi level (dashed
lines). The moment is equal to the difference of occupied t and t states and calculated
by integration over 0t (E) and OJ (E) (Skomski and Coey, 1999).

The Bloch character of itinerant wave functions means that the wave
functions extend to infinity. For two reasons, this is not real istic. First,
magnets encountered in real ity , in particular nanomagnets, cannot be
considered as infinite. Second, finite-temperature excitations create spin
disorder and break the Bloch symmetry of the t and ... wave functions. The
problem of nonequivalent sites can be tackled, for example, by real-space

<Z> In energy units, a field of 1 T (10 kOe) corresponds to /./0 /./e H = O. 058 meV. This
must be compared to single-electron energy differences of the order of 1 eV.
10 R.Skomski and D.J.Sellmyer

approaches (Cyrot-Lackmann, 1968; Heine, 1980; Desjonqueres and Spanjaard,


1993; Sutton, 1993; Skomski and Coey, 1999). When only nearest neighbors
are taken into account, these methods yield the correct band width, but details
of the band structure, such as peaks in the density of states, are ignored.
Increasing the number of neighbors improves the resolution of the density of
states and makes it possible to distinguish between bulk sites and sites close
to surfaces. A well-known numerical method to realize this scheme is the
recursion method (Heine, 1980; Desjonqueres and Spanjaard, 1993).

1. 2. 5 Magnetic Anisotropy

The energy of a magnetic solid depends on the orientation of the magnetization


with respect to the crystal axes, which is known as magnetic anisotropy.
Permanent magnets need a high magnetic anisotropy in order to keep the
magnetization in a desired direction. Soft magnets are characterized by a very
low anisotropy, whereas materials with intermediate anisotropies are used as
magnetic recording media. This section starts with the parameterization of the
problem and then discusses the physical origin of magnetic anisotropy.
It is convenient to express the magnetization M, for which IMI = M s ' as
M = Mss. The corresponding dimensionless and normalized magnetization
vector s is then

s = sinesin</>e + sinecos</>e + cosee z .


x y (1.2)

e,
In terms of </> and the simplest anisotropy-energy expression for a magnet of
volume V is then E a = K, Vsin 2 e. This anisotropy is known as lowest-order
(or second-order) uniaxial anisotropy, and K, is the first uniaxial anisotropy
constant. Since cos 2 e + sin 2 e = 1, the expression K 1 sin 2 e is equivalent to
e
- K, cos 2 = - K, s~. For arbitrary easy-axis directions n, as encountered for
example in polycrystalline materials, the expression K 1 sin 2 must be replacede
by- K, (n S)2. K, is widely used to describe uniaxial magnets (hexagonal,
tetragonal and rhombohedral crystals) and small ellipsoids of revolution (fine
particles). For K, >0 the easy magnetic direction is along the c (or z) axis,
which is called easy-axis anisotropy, whereas K, < 0 leads to easy-plane
anisotropy where the easy magnetic direction is anywhere in the a-b (or x-y)
plane.
For very low symmetry (orthorhombic, monoclinic and triclinic), the first-
order anisotropy energy is
Ea = K, V sin 2 e + K; V sin 2 e cos (2</, (1.3)

where K, and K,' are, in general, of comparable magnitude. This expression


must also be used for magnets having a low-symmetry shape, such as
ellipsoids with three unequal principal axes, and for a variety of surface
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 11

anisotropies, such as that of bcc (011) surfaces (Sander et al., 1996).


Higher-order anisotropy expressions contain, in general, both uniaxial and in-
plane terms. For example,

Ea
V
= K, sin 2 e+ K 2 sin 4 e + K; sin 4 e cos 44> (1. 4)

describes tetragonal, hexagonal, rhombohedral and cubic crystals. Hexagonal


and rhombohedral crystals are characterized by K; = 0 (fourth-order uniaxial
anisotropy), whereas in the tetragonal case K 2 and K; are of the same order
of magnitude.
In the case of cubic magnets, only two of the three anisotropy constants in
Eq. (1. 4) are independent. Cubic anisotropy is often written as

(1. 5)

where K~c) is the second cubic anisotropy constant. Analysis of Eq. (1. 5)
shows that K, >0 favors the alignment of the magnetization along the (001)
cube edges, which is called iron-type anisotropy. K, < 0 corresponds to an
alignment along the (111) cube diagonals and is referred to as nickel-type
anisotropy. Comparison of Eqs. (1. 3) and (1. 4) yields K 2 = - 7 K 1/8 +
K~c) /8 and K; = - K 1/8 + K~C) /8 (Skomski and Coey, 1999).
By definition, there are only even-order anisotropy terms. Odd-order
anisotropies may be caused by relativistic Moriya-Dzialoshinskii interactions,
exchange biasing, or particular micromagnetic regimes (Moorjani and Coey,
1984; Skomski et aI., 1998b). This refers in particular to unidirectional
anisotropies of the type K ud cos e,
which are observed as an asymmetry
(shift) of the hysteresis loop.
With respect to the physical origin of anisotropy it is necessary to
distinguish between magnetostatic effects and magnetocrystall ine anisotropy.
Figure 1. 6 illustrates that magnetostatic interactions give rise to shape
anisotropy. For homogeneously magnetized ellipsoids of revolution, the shape
anisotropy is given by

K I,sh = ~o (1 - 3D) M; (1.6)

where 0 = 0 z is the ellipsoid's demagnetizing factor (0 = 0 for long cylinders,


0= 1/3 for spheres, and 0 = 1 for plates). In noncubic crystals there is also a
magnetostatic contribution associated with dipole interactions between
neighboring atoms; this contribution is independent of the macroscopic shape
of the magnet.
The anisotropy of most materials reflects the competition between
electrostatic crystal-field interaction and spin-orbit coupling Al where s, s
and I are the spin and angular momentum operators, respectively, and A is
spin-orbit coupling constant. This anisotropy contribution was first considered
by Bloch and Gentile (Bloch and Gentile, 1931) and is known as
12 R.Skomski and D.J.Sellmyer

magnetocrystall ine anisotropy. The crystal field reflects the local symmetry of
the crystal or surface and acts on the orbits of the inner-shell d and f electrons.
For example, Fig. 1. 6d shows the charge density of a magnetic central atom in
a tetragonal environment of electron clouds. (The arrow describes the
direction of the spin of the central atom.) The shape and orientation of the
charge density depends on the crystal field and, via spin-orbit coupling, on the
spin orientation of the central atom. Changing the spin direction modifies the
electron cloud of the central atom and changes the crystal-field energy. This is
the source of magnetocrystalline anisotropy.

(a) (b) (c) (d)

Fig\lre 1. 6 Physical origin of magnetic anisotropy: (a - c) compass-needle analogy of


shape anisotropy and (d) magnetocrystalline anisotropy. From the point of view of
magnetostatic interaction, (a) is more favorable than (b), so that the easy magnetization
direction of a small elongated particle is parallel to the long axis. In macroscopic magnet
this mechanism is ineffective, because domain formation (c) interferes.

The magnitude of the anisotropy depends on the ratio of crystal-field


energy and spin-orbit coupling. Spin-orbit coupling is a relativistic effect, so
that }, is largest for inner electrons in heavy elements. In the case of 3d
atoms, the spin-orbit coupling }, ~ 50 meV is much smaller than the crystal-
field energy Eo;:?: 1 eV, and the magnetic anisotropy can be obtained
perturbatively. For uniaxial symmetry,
},2
Kl~- (1.7)
Eo
whereas the anisotropy of cubic materials scales as }, 4/ E6. In 3d oxides, Eo
is essentially equal to the electrostatic crystal-field splitting (Bloch and
Gentile, 1931), whereas in itinerant magnets it is roughly given by the width
of the d band (Brooks, 1940).
Rare-earth 4f electrons are close to the atomic core and exhibit a strong
spin-orbit interaction. This leads to a rigid coupling between spin and orbital
moment, and the magnetocrystalline anisotropy is given by the comparatively
small electrostatic interaction of the unquenched 4f charge clouds with the
crystal field ( Herbst, 1991; Coey, 1996; Skomski and Coey, 1999).
However, due to the absence of quenching effects, the interaction with the
crystal field is very effective, and rare-earth single-ion anisotropies are much
larger than typical 3d. This is exploited in advanced permanent magnets
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 13

(Section 1. 4. 1).
A variant of magnetocrystalline anisotropy is magnetoelastic anisotropy,
where crystal-field contributions are changed or created by mechanical strain.
For example, cubic magnets subjected to uniaxial stress exhibit some uniaxial
anisotropy. The magnetoelastic contribution to the first anisotropy constant is

(1. 8)

where a is the uniaxial stress and As is the saturation magnetostriction.


Experimental room-temperature values of A s are - 7 x 10 -6 T for iron, - 33 x
1O- 6 T for nickel, 40 x 1O- 6 T for Fe304' -1560 x 1O- 6 T for SmFez, 75 x 1O- 6 T
for FeCo, and practically zero for FezoNi so (permalloy). Another source of
magnetocrystalline anisotropy are surfaces and interfaces (Bland and Heinrich,
1994; Sander et al. , 1996; Neel, 1954; Millev et a!. , 1998).
Figure 1. 7 shows the temperature dependences of the anisotropy for some
ferromagnets. The main reason for the pronounced temperature dependence is
that typical anisotropy energies per atom are quite small, E af k B T ranging
from less than O. 1 K to a few K. The realization of room-temperature
anisotropy requires the support of the interatomic exchange field, which
suppresses the switching of individual atomic spins into states with reduced
anisotropy.

i 0.04
Fe

1
~

-{)'o2
~
~
Ni
;< 0.02
:.:: --{)04
.

ol..----!-_---.,.~I:L..,,-!-,o_ --{).06
o 800
--{).08L...-~,-----_~_-----o-'--c----
o 200
TC'C)

Co
i 0.4
1
~

~ 0.2 ~ 0
:.:: :.::
0
-20
-200 0 200 400 -200 0 400 600 800
TC'C) T('C)

Figure 1. 7 Temperature dependence of the magnetic anisotropy. Closed and open circles
are K 1 and K ic ) data. respectively.
14 R.Skomski and D.J.Sellmyer

1. 3 Extrinsic Properties

Magnetic properties derived from the hysteresis loop are extrinsic properties,
because they describe the real structure of the magnet. M-H hysteresis loops
are obtained by monitoring the volume-averaged magnetization M as a function
of the external magnetic field H, whereas B-H loops show the flux density B =
1J0 H + 1J0 M as a function of H. Figure 1. 8 shows typical M-H hysteresis
loops. Two very important extrinsic properties derived from M-H loops are the
coercive force or coercivity Hcand the remanent magnetization or remanence
M,. B-H loops are used, for example, to determine the energy product
(BH) max' which determines the maximum magnetostatic energy per magnet
volume stored outside the magnet. Table 1. 1 shows some intrinsic properties
of typical magnets. Hysteresis loops are usually corrected for the
demagnetizing field - DM, by plotting the magnetization as a function of the
internal field H-DM. This skewing or shearing correction makes the hysteresis
loops more rectangular. Virgin curves or initial curves are obtained on
increasing H from zero after thermal demagnetization, that is, after heating
beyond T c

Table 1. 1 Extrinsic properties of some typical magnets

PaM, Po He (BH)max
Material
(T) (T) (kJ/m 3 )

Cobalt steel (hard-magnetic) 10 0.025 8


Annealed iron (soft-magnetic) 1.0 0.0001 0.04
Sintered hexagonal ferrite 0.39 030 28
Anisotropic alnico 1.30 0.07 50
Metal-bonded SmCos 092 1.88 175
Polymer-bonded SmCos 0.58 1.00 60
Sintered Sm 2 Co1J!SmCoS 108 1.0 225
Sintered Nd-Fe-B 1.33 16 400
Polymer-bonded Nd-Fe-B 0.55 075 48

Hysteresis is a complex nonlinear, nonequilibrium and nonlocal phenomenon.


The determination of the local magnetization M(r), from which the hysteresis
loop is obtained by averaging, is further complicated by the influence of the
real structure (defect structure, morphology, metallurgical microstructure) of
the magnet. Hysteresis-related or extrinsic magnetic properties are also
known as micromagnetic properties (Brown, 1963a). This term is somewhat
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 15

-11'--:,1----.-1'

o -2 -I 0 2
J1oH(a.u.) J10H (T)
(a) (b)

Figure 1. 8 Magnetic hysteresis: (a) origin and phenomenology of hysteresis and


(b) typical M-H hysteresis loops (schematic). In (a), the dashed line starting at H = 0
and M = 0 is the initial or virgin curve. The insets in (a) are energy-landscape
equivalents of the magnetization.

unfortunate, because most micromagnetic phenomena, such as magnetic


hysteresis, are realized on submicron or nanostructural length scales.
Physically, hysteresis is a nonequilibrium phenomenon, reflecting the
existence of metastable energy minima: changing the applied field destroys
local minima, and the magnetization M (r) is forced to jump into another
minimum. The nonlocal character of extrinsic properties originates from the
involvement of magnetostatic dipole interactions; atomic-scale nonlocal
interactions, such as RKKY interactions, are of secondary importance on
length scales larger than about 1 nm (Skomski, 1999).
Micromagnetic problems are usually solved on a continuum level (Brown,
1963a; Aharoni, 1996; Skomski and Coey, 1999). For example, the
magnetization M s considered in micromagnetism is generally averaged over a
few interatomic distances and can be regarded as a temperature-dependent
but field-independent materials constant (micromagnetic parameter). Narrow-
wall phenomena, which have been studied for example in rare-earth cobalt
permanent magnets (Hilzinger and KronmQller, 1975) and at grain boundaries
(Skomski, 2001; Skomski et al. , 2001), involve individual atoms and atomic
planes and lead to comparatively small corrections to the extrinsic behavior.
In contrast to the intrinsic phenomena considered in Section 1. 2, which
affect the spontaneous magnetization M s = I M I, micromagnetic phenomena
are realized by local rotations of magnetization vector. The reason is that
typical micromagnetic energies are much smaller than the quantum-mechanical
energy contributions establishing Ms. To explain the hysteresis loop of
magnetic materials one needs to trace the local magnetization M(r) = M s s(r)
as a function of the applied field H. This is achieved by considering the free-
energy functional
16 R.Skomski and D.J.Sellmyer

F = f{ A"
M)J2
[ (M - K1
(nM)2
M; - 110 M H - 2110 M }
H d (M) d V.
5

(1.9)
As introduced in Section 1.2, M 5 (r) is the spontaneous magnetization, K 1 (r)
is the first uniaxial anisotropy constant, A (r) denotes the exchange stiffness,
and nCr) is the unit vector of the local anisotropy direction. H is the external
magnetic field, and Hdis the magnetostatic self-interaction field. The latter can
be written as Cl

Hd(r) = ~f 3(r- r')(r- 2


r'). M(r:) -I r - r' 1 M(r')dV'.
4rr I r - r 15
(1. 10)
The free-energy character of F reflects the intrinsic temperature dependence
of the parameters A, K 1 and M 5 (Section 1. 2). Furthermore, all parameters
entering Eq. (1. 9) are local parameters, because they depend on chemistry,
crystal structure, and crystall ite orientation. For example, the anisotropy
K 1 (r, n is easily changed by varying the chemical composition.

1. 3. 1 Coherent Rotation
In small particles, the exchange is sufficiently strong to ensure that M (r) is
constant throughout the magnet, that is, "M
in Eq. ( 1. 9) is zero. Depending
on the context, this regime is called coherent rotation (uniform rotation) or
Stoner-Wohlfarth reversal (Aharoni, 1962; Brown, 1963a). For uniaxial
ellipsoids of revolution having the symmetry axis parallel to the external field
=
H He z' the free energy is

~ = K 1 sin 2 e + ~0(1 - 30)M;sin 2 e -l1oM 5 Hcos e (1.11)

where 0 is the demagnetizing factor introduced in Section 1. 2. 5 and is the e


angle between M and ez . Figure 1. 8b shows this energy landscape for several
field values. The coercive field, at which the magnetization jumps from M =
M z = M 5 to M = - M 5 is obtained by stability analysis. Expanding Eq. (1. 11)
into small powers of yields e
~
V = (K 1+211(1-30)M5+2110M 5H)e
2 2
' ( 1. 12)

The free-energy minimum described by this equation vanishes when the

Cl This field differs by M/3 from the internal magnetostatic field obtained from
Maxwell's equations. However, magnetic fields couple as M H to the magnetization, so
that any term proportional to M M= M; amounts to a physically irrelevant shift of the zero-
point energy CSkomski and Coey, 1999b).
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 17

reverse field reaches the coercive field He = 2K 11 IJo M s + (1 - 3D) M s/2. This
Stoner-Wohlfarth coercivity is a simple example of a nucleation field
(Section 1. 3. 3).
It is often convenient to express anisotropies in terms of anisotropy fields.
For example, spherical Stoner-Wohlfarth particles (0 = 1/3) can be
described by H. = 2K 11 IJo Ms. Other widely-used anisotropy-field definitions
are based on the zero-field free-energy difference between parallel and
perpendicular orientations of the magnetization, H. = [F ( e = 1T 12) - F ( e =
0) JI IJo M s V, and on the saturation of the magnetization in a perpendicular
magnetic field. In general, the anisotropy field depends on its definition, but
with respect to lowest-order uniaxial anisotropy the different definitions all
reproduce the express ion 2K 1 1IJ 0 M s ( Skomsk i and Coey, 1999). H. is often
used as rough coercivity estimate, although coercivities encountered in
practice are often much smaller than the anisotropy field (Section 1. 3. 3) .
The Stoner-Wohlfarth approach can also be used to discuss the angular
dependence of hysteresis (Stoner and Wohlfarth, 1948). When the appl ied
field is parallel to the easy axis, as implied in Eqs. (1.11) and (1. 12), the
loop is rectangular and He = H. For fields perpendicular to the easy axis the
model yields a straight line M (H) with He = 0 and M ( H.) = Ms. For
intermediate field directions, He < H. but M ( H.) < Ms. The angular
dependence of the Stoner-Wohlfarth model is of some practical relevance,
because many magnetic materials can be approximated by nearly interaction-
free ensembles of uniaxial powder particles or polycrystallites
(nanocrystallites), corresponding to a superposition of Stoner-Wohlfarth
loops. For randomly oriented particles or grains, the remanence M r = M./2
and He=O. 479 H. By comparison, the remanence ratio Mr/M s equals 0.832
and O. 866 for iron-type (K 1 > 0) and nickel-type (K 1 < 0) cubic magnets,
respectively.
Adding fourth-order uniaxial anisotropy, that is, the K 2 term in Eq. (1. 4) ,
gives rise to a variety of zero-field spin configurations. When both K 1 and K 2
are positive, then the minimization of the anisotropy energy yields easy-axis
anisotropy (e = 0), whereas for K 1 < 0 and K 2 < 0 the magnetization lies in
the basal plane (easy-plane anisotropy, e = 1T 12). An interesting regime is
the easy-cone magnetism occurring if the conditions K 1 <0 and K 2 > - K 1/2
are satisfied simultaneously. The tilt angle between the z-axis and the easy
magnetization direction is given by

ee = arcsin.~
'\j ~ . (1. 13)

Since the temperature dependences of K 1 and K 2 are generally different (K 2 is


often negligible at high temperatures), the preferential magnetization direction
may change upon heating (spin-reorientation transition). A similar film-
thickness dependent transition is observed in films where surface and bulk
anisotropy contributions compete.
18 R.Skomski and D.J.Sellmyer

1. 3. 2 Domains and Domain Walls

The Stoner-Wohlfarth theory assumes that the magnetization is coherent


(uniform) throughout the magnet. This is justified for very small particles, but
in large particles or magnets the magnetostatic self-interaction leads to
magnetization inhomogenities. For example, magnetostatic interactions tend
to yield magnetic domains of opposite magnetization directions (Bloch, 1932;
Landau and Lifshitz, 1935; Kittel, 1949; Craik and Tebble, 1961). This
explains, for example, why spoons on a wedding table do not attract each
other, despite M(r)2 = M~ throughout the cutlery.
By expressing Eq. (1. 10) in terms of the magnetic charge density - VM
one can show that the self-interaction energy is lowered by the absence of
magnetic charges (poles) at the magnet's surface. Domains are separated by
comparatively thin domain walls (Bloch, 1932; Landau and Lifshitz, 1935;
Chikazumi, 1964). The reason is that the magnetization inside the domains
lies along easy directions, whereas the transition between two easy
magnetization directions involves energetically unfavorable spin orientations.
As a consequence, magnetocrystalline anisotropy favors narrow domain walls.
On the other hand, narrow walls correspond to large magnetization gradients
and are unfavorable from the point of view of exchange.
Figure 1. 9 shows some types of domains and domain walls. The
magnetization of bulk Bloch walls in uniaxial magnets obeys

Mz(x) =- M s tanh(x JK 1 /A) (1. 14)

where x is the distance from central plane of the wall, and A is exchange
stiffness. In thin films, as in Fig. 1. ge-f, the expressions for M(x) are more
compl icated due to magnetostatic interactions. Equation (1. 14) yields the
domain-wall width Ow = 2 ~, but in practice it is more common to base
the definition on the angle e
rather than on M z = M o cos so that 06 = e,
"IT ~ (Bozorth, 1951; Skomski and Coey, 1999). In both cases the
energy stored in the wall is given by the energy density y = 4 ~. Typical
domain-wall widths are 5 nm and 100 nm for hard and soft magnetic materials,

This example illustrates not only the distinction between soft and hard magnets but
also the distinction between soft magnets and paramagnets. High-alloy stainless steels are
paramagnetic at room temperature, but typical cutlery contains only small amounts of
anticorrosive additives such as Ni and Or and remains ferromagnetic. (Actually, at weddings
one expects silver spoons rather than steel spoons. There is anecdotal evidence that grooms
use small permanent magnets to check the financial of the future parents-in-law. If the spoon
is attracted by the magnet, then there is cheap steel under the silver coating. )
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 19

respectively.

~
~
(a) (b) (c) (d)

(0
Figure 1.9 Micromagnetic spin configurations: (a) single-domain state. as observed in
very small particles. (b) two-domain configuration. as encountered in fairly small particles
with uniaxial anisotropy. (c) flux-closure in cubic magnets. (d) complicated domain
structure in a polycrystalline magnet. (e) Bloch wall in a thin film with perpendicular
anisotropy. and (f) Neel wall in a thin film with in-plane anisotropy. In macroscopic
magnets, the width Ow of the domain walls separating the domains is much smaller than the
domain size. but in magnetic nanostructures the distinction between domains and domain
walls often fades.

The domain-wall width can also be estimated from dimensional arguments


(Bloch, 1932). The domain-wall width is determined by the anisotropy
constant K 1 and the exchange stiffness A, which are measured in J/m 3 and
J/ m, respectively, so that the only length and the only wall energy derivable
from these parameters are the wall-width parameter 00 = (A / K 1) 1/2 and the
wall-energy parameter Yo = (K 1 A) 1/2, respectively.
Magnetostatic self-interaction favors domain formation, but since the
creation of domain walls costs energy, there are no walls if the gain in
magnetostatic energy is smaller than the wall energy. For a wall separating
two semispherical domains in a spherical particle, the wall energy is YlTR 2 .
The competing gain in magnetostatic energy is roughly equal to half the single-
domain energy, that is J.lo M~ V/12, so that domain formation is favorable for
particles whose radius exceeds a critical single-domain radius

RSd~36~ (1. 15)


J.lo M s

This value varies between a few nm in soft magnets and about 1 IJm in hard
magnets. However, critical single-domain sizes - and domain sizes in
multidomain structures - are strongly geometry-dependent (Skomski et al. ,
1998c). For example, typical domains in films with perpendicular anisotropy,
as in Fig. 1. ge, often form meandering stripes. The domain size (stripe width)
o d is easily estimated by comparing the stray-field energy J.lo M~ 0 d L 2 , where
L 2 is the film area, with the wall energy ybL(L/D d ) . where b is the film
thickness (Kittel, 1949). Minimizing the total energy with respect to D d yields
20 R. Skomski and D. J. Sellmyer

Dd"",=,(yb/lJoM~)1/2. In other words, the film thickness exhibits a square-root


dependence on the film thickness.
It is important to keep in mind that the critical single-domain radius is
largely unrelated to hysteresis (Skomski and Coey, 1999). First, R sd is a
ground-state property, comparing free energies of single-domain and
multidomain states, whereas hysteresis is a nonequil ibrium phenomenon
caused by free-energy barriers. As we will see in the next subsection, the
onset of nonuniform (incoherent> reversal in perfect ellipsoids of revolution is
governed by the exchange length

lex = J110
AM2 .
s
(1. 16)

This quantity is anisotropy-independent and, in hard magnets, much smaller


than R sd (Table 1. 2). The popular but incorrect equating of single-domain
magnetism and coherent rotation has its origin in the focus on soft and semi-
hard magnets in the first half of the 20th century.

Table 1. 2 Micromagnetic parameters at room temperature


(The values for Fe and Ni are uniaxial estimates)

/JaMs A K1 0 Y I ex R Sd Ha
Material K
(T) (pJ/m) (MJ/m 3 ) (nm) (mJ/m 2 ) (nm) (nm) (T)

Fe 2.15 8.3 005 40 26 1.5 0.12 6 0.06


Co 1.76 10.3 053 14 9.3 2.0 0.46 34 0.76
Ni 0.61 3.4 -0 005 82 05 3.4 o 13 16 0.03
BaFe 12 019 0.47 6.1 0.33 14 5.7 59 1.37 290 1.8
SmCo5 1.07 22.0 17 36 77 49 4.35 764 40
Nd2 Fe14 B 1.61 7.7 49 39 25 1.9 1.54 107 7.6

The exchange length Eq. (1. 16) can be interpreted as the length below
which atomic exchange interactions dominate typical magnetostatic fields. It
also determines the thickness of soft-magnetic films below which Neel walls
are energetically more favorable than Bloch walls and the grain size of two-
phase magnets below which the hysteresis loops look single-phase like. From
an atom ic point of view, I ex is proportional to a 0/0' = 7 . 52 nm, where a 0 is the
Bohr length anc! 0' ""'=' 1/ 137 is Sommerfeld's fine-structure constant (Skomski
et al., 1998c). Note that the term exchange length is occasionally used to
denote the wall-width parameter 00 = (A / K I) 1/2. This is confusing, although
00 is important in a different context (localized nucleation, Section 1.3.3).
The hardness of a magnet is described by the dimensionless parameter
K = (K I / 110 M~) 1/2. (In magneto-optical recording,
2
K is known as the
"quality" parameter Q.) Magnetically very hard and very soft materials are
characterized by K 1 and Kl, respectively, as compared to lex""'='3 nm for
a broad range of magnetic materials (Skomski and Coey, 1999). In terms of
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 21

K, the critical single-domain radius and the wall-width parameter are equal to
36K/ ex and 'ex! K1/2, respectively. Table 1.2 shows typical micromagnetic
parameters.

1. 3. 3 Coercivity

The most intriguing aspect of hysteresis is the coercive force of coercivity. It


describes the stability of the remanent state and gives rise to the classification
of magnets into hard magnetic materials (permanent magnets), semihard
materials (storage media), and soft magnetic materials. A widely-used
phenomenological expression for the coercivity is

He = UK -2KM 1
- DeffM s -llH ( )
T,ry (1. 17)
1-10 s
where uKis the KronmOlier parameter (KronmOller, 1987; KronmOller et al.,
1988), D eff a magnetostatic interaction parameter, and llH is fluctuation-field
contribution caused by thermal activation (Neel 1951; Givord and Rossignol,
1996; Skomski and Coey, 1999). The comparatively smallllH term, which
will be discussed in Section 1. 5. 7, means that the coercivity depends on the
sweep rate ry=dH/dt.
Only in a few cases it is possible to obtain exact coercivity results. One
example is the coherent-rotation or Stoner-Wohlfarth model (Section 1. 3. 1).
However, it has been known for decades that neither the coercivity nor the
loop shapes of real materials are reproduced by the Stoner-Wohlfarth theory.
For example, the coercivity of optimized permanent magnets is only 20 % -
40 % of the anisotropy field. The main reason is imperfections, such as
metallurgical inhomogenities, grain boundaries, and surface irregularities.
Aside from mechanisms involving misaligned grains (Section 1. 3. 1 ) ,
there are two main coercivity mechanisms: nucleation and pinning.
Figure 1. 10 illustrates this difference. Nucleation-controlled magnets are
almost defect-free, and the coercivity is essentially given by the nucleation
field. There are several types of nucleation: coherent rotation in very small
particles, curling in large perfect ellipsoids of revolution, and localized
nucleation in imperfect structures (Skomski, 1998). In contrast to nucleation-
type magnets, pinning-type magnets contain many defects, which ensure
coercivity by impeding the motion of the domain walls.
Nucleation means that the original magnetization state (remanent state)
becomes unstable in a reverse magnetic field Hz = - H N The nucleation field
H N is often a good estimate for the coercivity, particularly in nearly perfect
magnets. Its determination of the nucleation field amounts to an eigenmode
analysis of the free energy Eq. (1. 9). It is convenient to write the local
magnetization as

(1.18)
22 R. Skomski and D. J. Sellmyer

/
,-
/
/
I
I

-11---:....------+--'

o
H (a.u.)

Figure 1. 10 Nucleation and pinning. (The dashed lines are virgin curves. ). In the case of
nucleation, the coercivity rei ies on the absence of domains. Once reversal has started in
nucleation-type magnets. the domain wall propagates nearly freely through the magnet.

where m is the perpendicular magnetization component. Inserting this


expression into Eq. ( 1. 9) and expanding the result into powers of m (r) yields
the quadratic free-energy expression

E = f[ A (\1 m)2 + Keff(r) m 2 - ~ 2


IJo M s Hm Jdr. (1. 19)

In this equation, K eff (r) is an effective local anisotropy field. For the coherent-
rotation and curl ing modes disCussed in this subsection. Eq. (1. 19) is exact.
in general there are corrections due to the non local character of the dipolar
self-interaction. Eigenmode analysis of Eq. (1. 19) yields the differential
equation

(1.20)

where we have assumed that A is constant throughout the magnet.


In structurally homogeneous ellipsoids of revolution, the coherent mode
mer) =mo is an eigenfunction of Eq. (1.20). The corresponding effective
anisotropy is K eff = K 1 + IJo (1 - 3D) M; /4, and the nucleation field is

HN = ~ + ~(1 - 3D)M s ( 121)


IJo Ms 2
This nucleation field is equal to the Stoner-Wohlfarth coercivity
(Section 1. 3. 1). Note that this result remains unchanged if one includes K 2
e
because sin 4 does not contain terms quadratic in e.
Coherent rotation is limited to very small particles. In large perfect
ellipsoids of revolution, nucleation is realized by the curling mode (Aharoni,
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 23

1962; Brown, 1963a; Aharoni, 1996). Figure 1. 11 compares the coherent-


rotation and curling modes. The curling nucleation field is

H N = ~ - OM + c (0) A (1.22)
fJ.oM s s fJ. o M s R 2

where the radius R = R x = R y refers to the two degenerate axes of the


ellipsoid, and c = 8. 666 for spheres (0 = 1/3) and c = 6.678 for needles
(0=0).
/-----,
~ -.- ---IJ

~~~1~
~~~1~
_---_/
.....

(a) (b) (c)

Figure 1. 11 Nucleation modes in homogeneous magnets: (a) coherent rotation in a


sphere, (b) curling in a sphere. and (c) curling in a cylinder. The arrows show the local
magnetization M= M z e z +m. where ez is parallel to the axis of revolution of the ellipsoid
(cylinder) .

Coherent rotation is favorable from the point of view of exchange, but the
exchange energy necessary to realize curling competes with the gain in
magnetostatic energy associated with the flux-closure clearly visible in
Fig. 1. 11. Comparison of Eqs. (1. 21) and (1. 22) yields a coherence radius
R c above which the nucleation occurs by curling. For homogeneous spheres
with uniaxial anisotropy one obtains R c = 5.099 lex' whereas in wires 2R c = 7.31
lex. Note that these radii are proportional to the exchange length and therefore
independent of the anisotropy constant K 1 (Section 1. 3. 2).
In hard magnets, where K 1 is large, the curl ing mode yields only small
corrections to the Stoner-Wohlfarth coercivity and cannot explain the observed
very low coercivities. The failure of this traditional theory is known as Brown's
paradox. Since the coherent-rotation and curling modes are exact solutions of
the nucleation problem in homogeneous ellipsoids of revolution, it is
necessary to consider deviations from the ideal of homogeneous ellipsoids of

As discussed for example in (Aharoni. 1996), other modes. such as those assumed
in (Braun and Bertram. 1994). lead to higher nucleation fields and are irrelevant to the
problem of ferromagnetic hysteresis. Thermal activation may. in principle. create a localized
nucleus of any energy. but due to the Boltzmann factor exp (- E a / k B T) the probability of
involving other modes is negligible (Aharoni. 1996; Sellmyer et al.. 2001).
24 R. Skomski and D. J. Sellmyer

revolution, that is, structural inhomogenities (Aharoni, 1962; Aharoni, 1996;


Brown, 1945; Kronmuller, 1987; Skomski et al., 1999b). In terms of
Eqs. (1.19)-(1.20), imperfections lead to a local reduction of the
anisotropy K 1 (r). This reduced anisotropy, as well as additional
demagnetizing-field and exchange corrections not considered in Eq. ( 1. 19) ,
yield a coercivity reduction (Skomski, 1992, 1998; Skomski et al., 1999b;
Zeng et al., 2002). For example, the nucleation field for a soft-magnetic
spherical inclusion of radius R in a very hard matrix corresponds to O'K = o~/
R 2 , where oBis the domain-wall width of the hard phase (Skomski, 1992).
The coercivity reduction is accompanied by a localization of the nucleation
mode. Localized nucleation corresponds to strongly inhomogeneous
magnetization states and is therefore unfavorable from the point of view of
interatomic exchange. By contrast, the coherent-rotation and curling modes
are delocal ized, that is, they extend throughout the magnet. On the other
hand, localization is favorable from the point of view of anisotropy, because it
exploits local K 1 (r) minima.
The pinning mechanism governs the magnetization reversal in strongly
inhomogeneous magnets and means that the coercivity is determined by the
interaction of domain walls with structural inhomogenities (Becker and Doring,
1939; Kersten, 1943). The trapping of walls by a small number of powerful
pinning centers is called strong pinning. A simple strong-pinning expression is
H p =[dy(x)/dx]/(2J../ o M s ) ' where y(x) is the average wall energy as a
function of the wall position. By contrast, pinning caused by a large number of
very small pinning centers, such as atomic defects, is called weak pinning. In
the case of weak pinning, the wall energy is averaged over a distance of order
OB' so that the density of pinning centers determines the pinning strength
(Gaunt, 1986).
One typical pinning mechanism involves inhomogenities whose anisotropy
constant is higher than that of the main phase: since high anisotropies yield
high domain-wall energies, the penetration of the wall into the high-anisotropy
regions is energetically unfavorable. This mechanism is known as repulsive
pinning, whereas the capturing of a wall in a low-anisotropy region is referred
to as attractive pinning. Figure 1. 12 illustrates the difference between
repulsive and attractive pinning. These mechanisms are realized, for
example, in pinning-type Sm-Co based magnets, where Sm2 (Co 1 - x T x ) 17
regions are separated by Sm (Co 1 - y T y ) 5 grain boundaries (T = Cu, Zr,
Ti, ... ) (Kumar, 1988; Skomski and Coey, 1999; Hadjipanayis, 1999; Zhou
et al., 2000). The pinning energy barrier is, in crude approximation,
proportional to the anisotropy difference K 1 (1 : 5) - K 1 (2 : 17), and by
changing the chemical composition or, for some compositions, the
temperature (Zhou et al. , 2000) it is possible to adjust the anisotropy and to
tune the pinning behavior.
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 25

1 1
1 1 Wall
I 1:5 1 motion
1 Phase 1
I" "I ~
f---+-----i_.I I._------i

Wall
(a) (b)
Figure 1. 12 Pinning at a grain boundary phase: (a) attractive pinning 8m-Co magnets and
(b) repulsive pinning. The energy landscape E(x) reflects the local chemistry.

1. 4 Magnetic Materials

As illustrated in Fig. 1.8b, a traditional criterion for the classification of


magnetic materials way of classifying magnetic materials is to consider their
magnetic "hardness" or coercivity. Historically, the term "hard" refers to the
comparison of pure iron and Fe-C alloys (steel): the addition of carbon
increases not only the mechanical hardness but also the coercivity. Other
important criteria are the chemical composition (oxides, transition-metal
elements, alloys) and the processing-related real structure (sintered, cast,
thin-film deposited, nanofabricated, etc.).

1. 4. 1 Permanent Magnets
Permanent magnets have a wide range of applications, for example in
electromotors, loudspeakers, windshield wipers, locks, microphones, and
hard-disk drives, and as toys and refrigerator magnets. Until the first half of
the 20th century, most permanent magnets were made from steel, and due to
the low coercivity of the magnets it was necessary to resort to cumbersome
horseshoe shapes. Steel magnets for permanent-magnet applications are now
obsolete, but the high saturation magnetization of Fe65 C0 35 and its pronounced
temperature stabil ity continue to be exploited in alnico magnets, whose
moderate coercivity originates from the shape anisotropy of elongated
Fe65 C035 particles embedded in a Ni-AI matrix.
Another material used in the past as a permanent magnet is magnetite
(Fe3 0 4 ). The anisotropy of this cubic ferrite is low, but hexagonal ferrites
such as BaFe12 0 19 exhibit an appreciable magnetocrystalline anisotropy and
are now widely used to produce cheap ceramic magnets with coercivities of up
to about O. 3 T (3 kOe). Magnetocrystalline anisotropy is also exploited in
26 R.Skomski and D.J.Sellmyer

tetragonal L1 0 magnets such as PtCo.


Most high performance magnets are now made from rare-earth transition-
metal intermetallics such as Nd2 Fel4 B (Sagawa et al., 1987) and SmCos
(Kumar, 1988), where the rare-earth sub lattice provides sufficient anisotropy
to realize broad hysteresis loops with coercivities of the order of 1 T (0.8 MAim).
These magnets consist of 3d atoms ensuring a high magnetization and a high
Curie temperature and rare-earth atoms ensuring a high uniaxial anisotropy
(Sagawa et al. , 1984, 1987; Croat et al. , 1984; Coey and Sun, 1990; Coey,
1991; Long and Grandjean, 1991; Coey and Skomski, 1993). As outlined in
Section 1. 2. 5, rare-earth anisotropy is caused by the interaction of aspherical
4f electron clouds with the anisotropic crystal field. The shape of the 4f
electron clouds follows from the Hund' s rules electronic structure of the
tripositive rare-earth ions and is described, in lowest order, by the Stevens
coefficient 0' J (Hutchings, 1964; Skomski and Coey, 1999). Some ions, such
as Nd3+ , are oblate, as in Figs. 1. 2b and 1. 6d, and characterized by O'J<O
whereas others, such as Sm3+, are prolate (0' J > 0). Since, for a given
crystalline environment, the sign of K I is given by the sign of O'J' it is possible
to tailor the anisotropy of rare-earth transition-metal intermetall ics by rare-
earth substitutions. In practice, coercivity as high as 4. 4 T has been achieved
in Sm3 Fel7 N3-based magnets (Kuhrt et al., 1992). A recent achievement in
the field is high-temperature Sm-Co-Cu-Ti alloys (Zhou et al. , 2000) having
coercivities above 1.2 T (12 kOe) at 500C .
A key figure of merit of permanent magnet is the energy product (BH) max'
which describes the ability to store magnetostatic energy. Energy product
increases with coercivity He and remanence M r but can never exceed the value
J.lo M~ I 4 ~ J.lo M~ 14. However, if magnetization were the only consideration
then 0' iron with J.lo M s = 2. 15 T would be used for permanent magnets with
energy products exceeding 920 kJ/m 3 . In fact, the coercivity of bcc iron is so
low that energy products of iron magnets are only of order 1 kJ/m 3 . However,
the coercivity problem has now been solved by the development of rare-earth
transition-metal permanent magnets, and high-performance Nd2 Fe14 B magnets
have room-temperature energy products as high as 405 kJ/m 3 . The means that
less than 109 of Nd2 Fe14 B are now able to replace a horseshoe steel magnet of
more than 2 kg.
On the other hand, rare-earth and non-magnetic elements reduce the
magnetization. The light rare-earths' atomic moments are at best slightly
larger than that of iron, but they occupy more than three times the volume.
Further progress in increasing the energy product therefore requires an
enhancement of the magnetization while ensuring that H e >M s /2 (Skomski and
Coeys 1993). One way of achieving this is nanostructuring (Section 1.5.6 and
Chapter 6).
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 27

1. 4. 2 Soft Magnets

Soft-magnetic materials are widely used for flux guidance in permanent-magnet


and other systems, in transformer cores and for high-frequency and microwave
applications and in recording heads. The main feature of soft magnets is their
low coercivity, He being several orders of magnitude smaller than in hard and
semihard materials. One figure of merit is the initial permeability IJr = (dM/
dH)o, which is small in permanent magnets and recording media but exceeds
1,000 in soft magnets (Evetts, 1992; McCurrie, 1994). High-frequency
applications require small hysteresis losses, so that the small coercivities are
often more important than a high permeability.
Iron-based metallic magnets have long been used as soft-magnetic
materials. Examples are pure iron, Fe-Si, Feso Coso, and permalloy (Ni so FE3:1o ). For
example, permalloy has an anisotropy of about O. 15kJ/m3 , an anisotropy field
of about O. 4 mT and a typical coercivity of about O. 04 mT (0. 4 Oe). More
recently, amorphous and nanostructured metals have attracted much attention
as soft-magnetic materials. Essentially, they have the composition T 100 - x Zx
(T=Fe, Co, Ni and Z=B, C, P, Zr, ... ) where x::=::::::10-20. Oxides, such as
simple ferrites (TFe2 0 4 , where T = Mn, Fe, Ni, Zn) and garnets
(R 3 FesO'2, R = Y, Gd, ... ) have a ferrimagnetic spin structure
(Section 1. 2. 3) and, therefore, a rather low magnetization. However, their
comparatively large resistively suppresses eddy-current losses and makes
them suitable for high-frequency applications, for example in antennas and
microwave devices (McCurrie, 1994).
Another important application of soft materials is inductive and
magnetoresistive recording heads (Comstock, 1999). The function of
inductive head materials is to realize flux closure for reading and writing on
recording media. Typical materials are Ni so Fe20 (permalloy, He = O. 01 to
0.05 mT), hot-pressed Ni-Zn and Mn-Zn ferrites (He = O. 02 mT); Fe-Si-AI
(sendust, He = 0.025 mT), as well as Fe-Ti-N and Fe-Rh-N alloys (Comstock,
1999). Magnetoresistive read heads exploit the anisotropic magnetoresistance
due to the spin-dependent scattering of conduction electrons by magnetic
atoms such as the central atom in Fig. 1. 6d or, more recently, the giant
magnetoresistive (GMR) effect exploiting the different Fermi-level spin-up and
spin-densities of the involved components (Fig. 1. 5). Almost all metallic
ferromagnets exhibit GMR, but soft-magnetic materials, such as permalloy,
are easier to switch. For the practical realization of GMR materials in
multi layers (spin valves) and granular materials see Section 1. 5. 1.
28 R.Skomski and D.J.Selimyer

1. 4. 3 Recording Media

Magnetic recording media are of great importance in computer and audio-


visual technology. They are used, for example, in magnetic tapes and for data
storage in hard-disks (Velu and Lambeth, 1991; Sellmyer et al.. 1998;
Comstock. 1999; Weller et al. , 2000). Semihard magnetic materials used in
storage media exhibit more or less rectangular hysteresis loops having
coercivities of the order of O. 1 T (80 kA/m). The coercivity of storage media
is sufficient to assure the remanence of the stored information without requiring
powerful and bulky writing facilities. The rectangular character of the
hysteresis loops ensures a well-defined switching behavior; it is described by
squareness parameters such as the remanence ratio S' = M r / M s and the
coercive loop squareness S' =1-(Mr/H e)/Xe' where Xe=dM/dH(H e ).
Traditional storage media are made using materials such as granular
Fe203 and Cr02' The disadvantage of those materials is the comparatively low
anisotropy. Advanced high-density recording media. characterized by more
than 10 Gb/ in 2 ( 1. 55 Gb/cm 2 ). are based on materials such as Co-Cr-Pt-B,
where the Pt improves the anisotropy. Other classes of materials, such as
rare-earth transition-metal nanocomposite films (Sellmyer et al. , 2002; Velu
and Lambeth, 1991) are also being considered. Some limitations in magnetic
recording are the present use of longitudinal or in-plane recording media. with
spin configurations similar to Fig. 1. 9f, and thermal instabilities at ultrahigh
recording densities (up to about 100 Gb/ in2 ) . These problems can be tackled
by suitable nanostructuring. using perpendicular recording (Wood, 2000)
similar to Fig. 1. ge and highly anisotropic media (Weller et al., 2000) to
avoid thermal demagnetization (Section 1.5.7).
A related field is magneto-optical recording using thin-film materials such
as amorphous Tb 22 Fe66 C0 12 (Comstock. 1999). The moderate anisotropy of
these many-sub lattice materials is magnetoelastic. caused by growth-induced
strain. However. since He """ 2K 1/110 Ms. the low magnetization near a
compensation point, as shown in Fig. 1. 4, ensures a sufficiently high
coercivity. On the other hand. the low magnetization has no negative impact
on the magneto-optical reading of the information, because the compensation
involves anti parallel transition-metal and rare-earth sub lattices having equal
magnetizations but different Kerr-response characteristics. The magnetic and
magneto-optical properties of the films can be tuned by choosing suitable rare-
earth and transition-metal stoichiometries, and the information can be stored
by thermomagnetic and Curie-temperature writing involving T k and T e
respectively (Comstock. 1999).
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 29

1. 4. 4 Other Magnetic Materials

Aside from the above-mentioned main groups of magnetic materials, there are
several types of magnetic materials for special present or future applications.
A topic of considerable current interest is spin-electronics (Ziese and
Thornton, 2001). The idea is to exploit the spin as an additional degree of
freedom in various types of electronic devices. Examples are giant
magnetoresistive devices consisting of multilayered and granular materials,
rare-earth and transition-metal based magnetic semiconductors, and materials
for distant-future quantum computing. In spin-electronics, the distinction
between bulk materials, nanoparticles, and devices blurs; some aspects of
this will be discussed in Section 1. 5.
Small magnetic Stoner-Wohlfarth particles are used not only as so-called
"elongated single-domain particles" (Dormann et al., 1994) in magnetic
recording but also in stable colloidal suspensions known as ferrofluids (Charles,
1992). A variety of materials can be used, such as Fe3 0 4 , BaFe12 0 19 , Fe, Co,
and Ni, and a typical particle size is 10 nm. Most ferrofluids are based on
hydrocarbons or other organic liquids, whereas water-based ferrofluids are
more difficult to produce. A characteristic feature of the magnetization
dynamics of ferrofluid particles is the distinction between Brownian relaxation
and Neel relaxation. Neel relaxation involves jumps over magnetic energy
barriers (Section 1. 5. 7) and quantum tunneling (chapter 5), whereas
Brownian relaxation reflects from the mechanical rotation of the particles due
to Zeeman interaction. The Brownian relaxation time is Ts = 3 V1)/ k s T, where
1) is the mechanical viscosity of the embedding liquid. Ferrofluids are used as
liquids in bearings and to monitor magnetic fields and domain configurations.

1. 4. 5 Tables

Tables 1. 3 and 1. 4 show the magnetic moment m, the spontaneous


magnetization M s' the Curie temperature T c' and first uniaxial anisotropy
constant K 1for some magnetic materials (Skomski and Coey, 1999; Evetts,
1992). Not included are antiferromagnets, such as NiO, GdFe03' and Ti 20 3.
The metastable compound y-Fe2 0 3has a moment of 2. 5 /-Is per formula unit
and a Curie temperature of 950 K, but above about 400 'c it transforms into
o:-Fe203 (McCurrie, 1994).
30 R.Skomski and D.J.Sellmyer

Table 1. 3 Intrinsic properties of some transition-metal oxides and alloys.


M s and K, are room-temperature values.
Substance lJoMsCT) Te(K) K,(MJ/m 3 ) Structure

Fe 2. 15 1043 o. 048 cubic bcc


Co 176 1388 0 53 hex. hcp
Ni 0.62 631 - 0.0048 cubic fcc
PtCo 1.00 840 4.9 tetr. CuAu (I)
PtFe 1.43 750 6.6 tetr. CuAu (I)
PdFe 1.37 760 1.8 tetr. CuAu (I)
MnAI 0.62 650 1.7 tetr. CuAu (I)
MnBi 0.78 630 1.2 hex. NiAs
Fe 3 0, O. 60 858 - O. 011 cubic MgAbO,
MnFe20, 0.52 573 - 0.0028 cubic MgAbO,
CoFe20, 0.50 793 0.270 cubic MgAbO,
NiFe20, O. 34 858 - o. 0069 cubic MgAbO,
CuFe20, 0.17 728 -0.0060 cubic MgAbO,
MgFe20, O. 14 713 - O. 0039 cubic MgAbO,
BaFe'20'9 0.48 723 0.330 hex. M ferrite
SrFe'20'9 0.46 733 0 35 hex. M ferrite
PbFe'2 0 '9 0.40 724 0.22 hex. M ferrite
BaZnFe'7 0 27 O. 48 703 O. 021 hex. W ferrite

Y3 Fes O'2 0.16 560 -0.00067 cubic ( garnet)


Sm3FesO'2 0.17 578 -0.0025 cubic (garnet)
DY3Fes0 12 O. 05 563 - O. 0005 cubic (garnet)
Cr02 o 56 390 0 025 tetr. Ti0 2
NiMn03 o 13 437 - 0.26 hex. FeTi03
y-Fe 2 03 0.47 863 -0.0046 cubic disordered spinel

Table 1. 4 Intrinsic properties of some rare-earth transition-metal intermetallics.


M s and K, are room-temperature values.

Substance K, (MJ/m 3 ) Structure

NdCos 1.23 910 0.7 hex. CaCus


SmCos 107 1003 17.0 hex. CaCus
YCos 1.06 987 5.2 hex. CaCus
Pr2 Fe" B 1.41 565 5.6 tetr. Nd2Fe" B
Nd2Fe"B 1.61 585 50 tetr. Nd2Fe" B
Sm2Fe"B 1.49 618 -12.0 tetr. Nd2Fe" B
DY2 Fe "B 067 593 4.5 tetr. Nd2Fe" B
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 31

continued
3
Substance fJo M s (T) T c CK) K 1 CMJ/m ) Structure

Er2 Fe14 B 0.95 557 -0.03 tetr. Nd2Fe14 B


Y2Fe14 B 1.36 571 1.06 tetr. Nd2Fe 14 B
SmCFell Tj) 1.14 584 4.9 tetr. ThMn12
YCFell Tj) 1.12 524 0.89 tetr. ThMn12
YCCO ll Tj) 0.93 940 -0.47 tetr. ThMn12
Nd2C0 17 1.39 1150 -1.1 rhomb. Th2Zn\7
Sm2 CO\7 1.20 1190 3.3 rhomb. Th2Zn17
DY2 CO \7 0.68 1152 -2.6 rhomb. or hex. Th2Zn\7 or Th 2Ni\7
Er 2Co\7 0.91 1186 0.72 hex. Th2Ni 17
Y2CO l7 1.25 1167 -0.34 rhomb. or hex. Th2Zn\7 or Th2 Ni 17
Sm2 Fe\7 1.17 389 -0.8 rhomb. Th2Zn17
Sm2 Fe\7N 3 1.54 749 8.9 rhomb. Th2Zn17
Y2Fe\7 0.84 320 -0.4 hex. Th2Ni\7
Y2Fe\7 N3 1.46 694 -1.1 hex. Th2Ni\7

1. 5 Magnetic Nanostructures

The magnetism of nanostructures exhibits various scientifically interesting and


technologically important deviations from that of magnetic bulk compounds.
First, the intrinsic behavior of magnetic nanostructures is modified by finite-
size and surface effects, and one may wonder how many atoms are necessary
to approach the intrinsic behavior of bulk magnets (Section 1. 5. 2). Second,
the structures exhibit a rich intrinsic behavior, including phenomena such as
random-anisotropy scaling (Chudnovsky et ai., 1986 ), remanence enhancement
(Coehoorn et al., 1988), micromagnetic localization (Skomski, 1998a),
bulging-type nucleation modes (Skomski et a!., 1999b), and a variety of
exchange-coupling (Hadjipanayis, 1999; Liu et al. ,2000) and grain-boundary
effects (Skomski et al., 1998b).
Some nanostructures exist in nature, or can comparatively easily be
produced in the bulk (Fig. 1.13) or in thin films. An example of natural
nanomagnetism is the magnetotaxy of bacteria living in dark environments
CCraik, 1995). Magnetotactic bacteria contain chains of particles of magnetite
(Fe 3 0 4 ) , having sizes of the order of 40 to 100 nm and used for vertical
orientation. Similar magnetite particles have been found in the brains of bees,
pigeons, and tuna, and it is being investigated whether they contribute to bird
migration. Some nanostructured materials, such as ferrofluids and sintered
32 R. Skomski and D. J. Sellmyer

Sm-Co and alnico permanent magnets, are produced by comparatively simple


techniques, and their further development and refinement is a challenge to
present-day technology. For example, nanoparticle ferrofluids are being
considered for cancer treatment, guided by a magnet and delivering high local
doses of drugs or radiation.

(a) (b)

Figure 1. 13 Two schematic bulk nanostructures: (a) sintered 8m-Co and


(b) hypothetical magnetic clusters (white) embedded in matrix. The two structures are
very different from the point of view of size, geometry, origin, and functionality. The Sm-
Co magnets, consisting of a rhombohedral Sm,Cop-type main phase (gray), a Cu-rich
SmC0 5 -type grain-boundary phase (black), and a Zr-rich hexagonal Sm, COl? -type
platelet phase (white), are produced by a complicated annealing process and widely
used in permanent magnets. Nanostructures such as that shown in (b) can be produced,
for example, by mechanical alloying and are used as permanent magnets, soft magnets,
and magnetoresistive materials (Berkowitz et al. , 1992; Xiao et al., 1992).

However, the most fascinating aspect of advanced magnetism is artificial


nanostructuring, embracing a materials-by-design concept. This "meta-
materials" approach makes it possible to produce materials not encountered in
nature. One example is hard-soft nanocomposites (Coehoorn et al., 1988,
Skomski and Coey, 1993). By suitably nanostructuring it is possible to create
two-phase magnets whose performance goes beyond what is expected from
the volume fractions of the involved phases. In other words, the addition of a
soft phase actually improves the permanent-magnet performance of the
material (Skomski and Coey, 1993; Liu et al., 1998).

1. 5.1 Physical Classification of Magnetic Nanostructures

Advanced magnetic nanostructures are characterized by a fascinating diversity


of geometries, ranging from intriguingly complex bulk structures to a broad
variety of low-dimensional systems. A comprehensive description of the
fabrication and characterization of specific nanostructures (Sellmyer et al. ,
2002; Himpsel et al., 1998; Gradmann, 1993, Zangari and Lambeth, 1997;
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 33

Sellmyer et al. , 2001) goes far beyond the scope of this chapter, but it is
appropriate to summarize some recent developments in this field.
Magnetic thin films and multi layers have attracted much interest in the
context of nanomagnetism, although they are often considered as a separate
branch of condensed-matter physics (Baibich et a!. , 1988). Among the many
highly interesting features are surface anisotropies (Neel, 1954; Millev et a!. ,
1998; Gradmann, 1993), moment modifications at surfaces and interfaces
(Bland and Heinrich, 1994), thickness-dependent magnetostrictive and
coercive phenomena (Sander et al., 1996; Oepen and Kirschner, 1989),
interlayer exchange-coupling (Baibich et a!., 1988), and finite-temperature
magnetic ordering (Bander and Mills, 1988). Two specific phenomena are the
nanoscale exchange-coupling or "exchange-spring" effects in multi layers
(Kneller and Hawig, 1991; Nieber and KronmOller, 1989; Skomski and Coey,
1993; Bowden et a!. , 2000; Sawicki et a!. , 2000; Parhofer et a!., 1996;
Fullerton et a!., 1999), and the pinning of domain walls in sesquilayer iron-
tungsten film structures (Sander et a!. , 1996).
It is virtually impossible to overestimate the importance of magnetic dots,
dot arrays, and patterned media. As mentioned in Section 1. 4. 3, advanced
magnetic recording media can be characterized as a complex array of
magnetic particles, and the call for well-characterized large-area arrays of
nanoparticles has stimulated the search for advanced production methods such
as laser-interference lithography (L1L), where laser-intensity maxima effect a
local decomposition of a nonferromagnetic material into ferromagnetic islands
(Zheng et a!. , 2001). A traditional though cumbersome method to produce
patterned arrays is e-beam nanol ithography (Chou et al. , 1994; Hehn et al. ,
1995). Some other production methods are discussed in (Himpsel et al. ,
1998; Nielsch et al., 2001). A recent example of an artificial fabrication
method is focused ion-beam milling (FIB), used to create small particles and
particle arrays with well-defined properties (Warin et al. , 2001). In a general
sense, this category of nanostructures also includes antidots, that is, dots of
nonmagnetic materials in a magnetic thin film.
There is a smooth transition from elongated dots and thin-film patches
(Shen et a!., 1997; Wirth et a!., 1998) to nanowires. A vehicle for the
fabrication of magnetic nanostructures (Roxlo et al., 1987; Masuda and
Fukuda, 1995; Sugawara et al., 1998; Sellmyer et a!., 2001) is the
deposition into molecular sieves (Ozin, 1992), track-etched polymer
membranes (Martin, 1994; Fert and Piraux, 1999) and porous anodic alumina
(Routkevitch et a!. , 1996; Sugawara et a!. , 1998). By electrodeposition into
alumina it is now possible to produce Fe, Co, and Ni wires with diameters of
about 5 nm and lengths of the order of 1 IJm, and on a somewhat larger scale
but without additional measures one can assemble nearly hexagonal nanowire
arrays with variable center-to-center spacings of the order of 50 nm (Sellmyer
et al. , 2001; L i and Metzger, 1997; Jessensky et al., 1998; Zeng et al. ,
2000). The resulting materials might be used, for example, as magnetic
34 R. Skomski and D. J. Sellmyer

recording media (Kawai and Ueda, 1975; Shiraki et al., 1985), optical
devices (Saito et al. , 1989), and electroluminescent display devices (Mizuki
et al. , 1987). By comparison, typical nanowires deposited in polymers, such
as Ni wires deposited in porous polycarbonate membranes (Wernsdorfer
et al. , 1997), have diameters between 20 and 200 nm.
As mentioned above, embedded clusters, granular materials, and other
bulk nanostructures are of great importance in nanoscience. Figure 1.13
presents two schematic examples of bulk nanostructures. Figure 1. 13a shows
a pinning-type Sm-Co magnet, produced by a complicated heat-treatment
procedure (Kumar, 1988). Structures similar to Fig. 1. 13b can be produced
by methods such as mechanical alloying; depending on grain size and
microchemistry, they are used, for example, as permanent magnets (Nd-Fe-B),
soft magnets (Fe-Cu-Nb-Si-B), and magnetoresistive materials (Co-Ag).
There are two types of exchange-coupled permanent magnets. Isotropic
exchange-coupled magnets CCoehoorn et al. , 1988; Muller et al. , 1991; Ding
et al., 1993; Manaf et al., 1993; Lesl ie-Pelecky and Schalek, 1999) are
random-anisotropy nanostructures and exploit remanence enhancement
(Section 1.5.6), whereas oriented hard-soft composites (Liu et al., 1998;
Parhofer et ai., 1996) utilize exchange coupling of a soft phase with a high
magnetization to a hard skeleton (see Chapter 6). A closely related system
with many potential appl ications is magnetic clusters deposited in a
nanomagnetic matrix. For example, the narrow size distribution of 10% -20%
makes these materials interesting as a granular media for magnetic recording
(Sellmyer et al. , 2002).
Finally, miniaturization leads to the development of nanodevices. This
category, which is only briefly mentioned here, includes for example
magnetoresistive spin valves, nanojunctions, micro-electromechanical systems
(MEMS), and magnetic force microscopy (MFM) tips.

1. 5. 2 Intrinsic Properties and Finite-Size Effects


Surface and interface atoms in nanostructures yield a disproportionally strong
contribution to the intrinsic behavior of the material. Figure 1. 14 shows the
modification of the moment and of the effective interatomic exchange in
multilayered Pt-Fe magnets (Sabiryanov and Jaswal, 1998a). Aside from the
strong moment fluctuations reflect the layered structure of L1 0 PtFe, the figure
shows effects extending over a few interatomic distances. These features may
be considered as a band-structure analog to the RKKY interaction of local
moments in a free electron gas. However, the total contribution of these
contributions is typically very small on length scales of more than 1 nm.
Concerning the finite-temperature spontaneous magnetization and the
Curie temperature, ordered nanostructures exhibit finite-size and surface
corrections well-known from thin-film and surface science (Bander and Mills,
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 35

Pt-Fe Fe Pt-Fe Fe > 200 Pt-Fe Fe Pt-Fe Fe

"'2>'"
<l)

-5...
E
<l) ~
2
E oj
0-
100
u <l)
.~ OJ)
(:: Fe
~ * Pt .J::
u
:2 ><
w
0 0
Atomic layer Fe-Atom layers
(a) (b)

Figure 1. 14 Intrinsic properties of multilayered Pt-Fe structures (Sabiryanov and


Jaswal 1998a).

1988; Maciejewski and Duda, 1987; Skomski et ai., 1998d). In disordered


nanostructures, the problem can be mapped onto a multi-sublattice problem
(Skomski and Sellmyer, 2000). Disordered two-phase nanostructures have a
single common Curie temperature close to the Curie temperature of the phase
with the strongest exchange coupl ing. However, when the grain size is larger
than a few interatomic distances, then the M s (T) curve of a inhomogeneous
ferromagnet is difficult to distinguish from a superposition of different phases.
The reason is the perturbations decay as exp[ - R / R o (T) J, where R o is the
"local" correlation length (Skomski and Sellmyer, 2000). This example shows
that intrinsic properties are realized on fairly small length scales, even if the
range of critical fluctuations goes to infinity.
Since Neel' s pioneering work on magnetic surface anisotropy ( Neel,
1954), the important role of this contribution has been recognized. This refers
in particular to complicated structures and morphologies such as ultrathin
transition-metal films (Gradmann, 1993), multi layers (Victora and McLaren,
1993), rough surfaces (Johnson et al. , 1996), and surface steps (Johnson et
al. , 1996; Chuang et al. , 1994), where first-principle calculations (Gay and
Richter, 1986; Bruno, 1989; Daalderop et ai., 1990; Victora and McLaren,
1993; Wang et ai., 1993; Bland and Heinrich, 1994) have been used to
calculate the anisotropy. Figure 1. 15 illustrates the atomic origin of the
surface anisotropy. An important point is that surface anisotropies easily
dominate the bulk anisotropy of cubic materials. From Tables 1.3 and 1. 4 we
see that bulk anisotropies are about two orders of magnitude smaller than
lowest-order (uniaxial) anisotropies. Due to the comparatively large number
Nsof surface atoms of particles (clusters), the surface contribution dominates
the bulk anisotropy in particles smaller than about 3 nm, even if one takes into
account that the net surface anisotropy is not necessarily linear in N s but tends
to scale as N~/2 due to random-anisotropy effects (compare Section 1. 5. 6) .
Note that this anisotropy is generally biaxial, that is, it contains both K 1 and
K; (Section 1. 2. 5).
36 R.Skomski and D. J. Sellmyer

X
bee (00 I)
~
bee(OO I)
>K ~
bee(Oll) bee(lll)

Figure 1. 15 Surface anisotropy and symmetry. In spite of the cubic character of bcc
magnets, (011) and (111) surfaces give rise to second-order and sixth-order in-plane
anisotropy contributions, respectively.

1. 5. 3 Narrow-Wall and Constricted-Wall Phenomena

Typical domain walls are smooth and extend over many interatomic distances
(Table 1. 2). However, deviations from this continuum picture occur in very
hard materials (narrow walls), at grain boundaries, and in the case of
geometrical constraints. Narrow domain walls in e,xtremely hard materials,
such as SmCos , call for the consideration of individual atomic planes but yield
only small corrections to the magnetic properties (Hilzinger and Kronmuller,
1975), The grain-boundary structures are of interest in several areas, For
example, they affect the spin-dependent scattering at nanojunctions (Ziese and
Thornton, 2001; Versluijs et al. , 2001) and the hysteresis loops of granular
permanent magnets (Skomski et al. , 1998a),
The importance of grain boundaries has been known for many years
(Fukunaga and Inoue, 1992), but from a micromagnetic point of view grain
boundaries and geometrical constraints were first considered by Skomski
et al. (1998a, 2001; Skomski, 2001), Figure 1, 16 shows two typical
examples, A particular feature of these spin structures is a quasi-
inhomogeneity at interfaces and small junctions (Skomski, 2001; Skomski et
ai, , 2001). Note that the effective exchange between grains is typically much
smaller than the value N a J, where J is the interatomic exchange (Section
1. 2, 3) and N a is the number of pairs of adjacent interfaces,
Concerning the extrinsic properties, a key theoretical problem is to
derive magnetization curves by simulation or modeling (Skomski and Coey,
1993) the magnet's nanostructure, Much progress has been made in the field
of numerical simulation (Schrefl et al., 1994; Schrefl and Fidler, 1998;
Fischer and Kronmuller, 1998), but some problems are more conveniently
tackled by analytic calculations. On the one hand, the size of the simulated
structures is limited to about 100 x 100 x 100 lattice points. In practice, this
excludes phenomena such as layer-resolved interatomic exchange on length
scales smaller than 1 nm and the formation of equilibrium domains on length
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 37

j j

j j j

/" '\
"- I)
~

(b)

Figure 1. 16 Exchange effects at interfaces: (a) reduced interface exchange and


(b) narrow junction.

scales larger than 100 nm. On the other hand, it is sometimes difficult to map
large amounts of numerical data onto transparent physical ideas (Binder,
1999).

1. 5. 4 Nanomagnetic Localization

As mentioned in Section 1. 3. 3, defects lead to local ization of the nucleation-


and more generally - reversal mode, a reduction of the coercivity. As one may
deduce for example from thermal activation volumes (Section 1.5.7), typical
local ization lengths range from less than 5 nm to more than 20 nm. In this
subsection, we will study two examples: clamped nucleation modes in aligned
two-phase nanostructures and one-dimensional Anderson local ization in thin
nanowires.
Clamped modes are important in the context of two-phase nanostructures.
Figure 1. 17 shows the geometry for a spherical soft magnetic inclusion (white)
in an aligned hard matrix (gray). Several cases have been considered in the
past, such as a simple solution of the problem for small particles in an infinitely
hard matrix (Skomski, 1992), small inclusions in a matrix of arbitrary
anisotropy and exchange stiffness (Skomski and Coey, 1993), and the exact
solution for particles of arbitrary diameter in a very hard matrix (Skomski
et aI., 1999b). Note that similar localized modes exist in some
inhomogeneous core-shell, nanowire, and thin-film geometries (Skomski
38 R. Skomski and D. J. Sellmyer

et al. , 1999b), and loosely related qual itative (Kneller and Hawig, 1991) and
quantitative (Nieber and Kronmuller, 1989; Skomski and Coey, 1993;
Skomski, 1998) considerations are impl ied in the context of exchange-spring
multi layers.

(a) (b)

Figure 1. 17 Nucleation modes in spheres surrounded by a hard-magnetic shell:


(a) bulging and (b) clamped curling. The figures are top views on the equator
plane. and the arrows show mer) for the core phase (white); in the hard-magnetic
environment (gray) m(r)~O. In both cases. the radial dependence of m is given by
spherical Bessel functions and localized in the soft region.

The starting point of the of the calculation of the modes shown in Fig. 1. 17
is the energy functional Eq. ( 1. 9). The eigenmode analysis-a generalization
of Eqs. (1. 11) and (1. 12 )-yields a bulging mode, characterized by the
symmetry of the coherent-rotation mode but incoherent due to the radial
dependence of m, and a clamped curling mode (Skomski et al., 1999b).
Figure 1. 17a, b show the spin structures of the quasi-coherent bulging and
clamped curling modes respectively; clamped curling is realized for large
inclusion diameters.
As a second example, we consider thin transition-metal nanowires
deposited in alumina (Section 1. 5. 1). The leading anisotropy contribution in
soft and semihard wires is shape anisotropy (Huysmans and Lodder, 1988;
Fert and Piraux, 1999; Sellmyer et al., 2001). By comparison, the bulk
magnetocrystalline anisotropy of Fe, Ni and fcc Co is small, because the
crystal symmetry is cub ic, and its influence is further reduced by the
polycrystalline character of typical wires (Zheng et al., 2000; Sellmyer et
al., 2001). The magnetostatic interaction between the wires may be
approximated by a comparatively small mean-field-type Zeeman field (Zeng et
al. , 2000; Sellmyer et al. , 2001) for the considered thin wires R < R c' so
that magnetization reversal in perfect wires starts with coherent rotation rather
than with curling (Zheng et al. , 2000; Sellmyer et al. , 2001; Skomski et al. ,
2000). However, as in bulk magnets, the coherent-rotation or Stoner-
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 39

Wohlfarth theory overestimates the coercivity (Section 1. 3. 3) and the


activation volume (Section 1. 5. 7 ). This disagreement is explained by
nanomagnetic (or micromagnetic) localization (Sellmyer et aI., 2001).
Applying Eq. (1. 9) to a nearly perfect thin cylinder yields, after some
calculation, the energy expression (Skomski et al. , 2000)

(1.23)

where K eff = K u (z) + M~ /2 is the effective anisotropy of the wire, 4> denotes
the angle between the wire axis (e z ), and H = Hz is the external magnetic
field. The inhomogenity consists in a local reduction of the anisotropy due to
wire-thickness fluctuations or chemical inhomogenities. For simplicity we
assume that K eff (x) = K 0 - L ilKS (x), where ilK is the magnitude of the
reduction and L is its extension along the z axis. As in Section 1. 3. 3, the
nucleation mode 4> (z) is obtained by minimizing Eq. (1. 23) and solving the
resulting eigenvalue problem. Away from the perturbation, the mode decays
exponentially, 4> (z) ....... exp( - z / R L ) , where the localization length is given by
(Skomski et al. , 2000)
2A
RL = --,--,--,------,--,--,-.,-
(K - Ks)L .
(1. 24)
o
The localization of the nucleation mode is accompanied by the coercivity
reduction:

(1.25)

Equation (1. 24) shows that degree of local ization (the local ization length) is
strongly dependent on the wire's nanostructure. For zero disorder, the
localization length goes to infinity and the reversal degenerates into coherent
rotation, as expected for ideal nanowires.
According to Eq. (1. 24 ), arbitrarily small disorder gives rise to
localization. Exploiting the analogy between micromagnetism and quantum
mechanics (Skomski and Coey, 1993) we can understand this finding in terms
of the localization behavior of a one-dimensional electron gas (Anderson,
1958). On the other hand, Fig. 1. 18 indicates that the radial dependence of
the coercivity obeys the expected (Brown, 1963a; Skomski and Coey, 1999;
Ross et al. , 2000) transition from coherent rotation to curling (Zeng et al. ,
2002) .

In the present context, it is of secondary importance whether the localization


originates from polycrystallinity, wire-thickness fluctuations, defects, impurities, or
geometrical features at the wire ends, or from a combination of these factors.
40 R.Skomski and D. J. Sellmyer

4000

3500

3000
<lJ
0 2500
bu
::r::
2000

1500

1000
0 10 20 30
R(mm)

Figure 1. 18 Dependence of the coercivity on the wire radius of Fe (top) and Co (bottom)
wires deposited in alumina (Zeng et al. , 2002). The localization essentially occurs along
the wire axis; the radius dependence exhibits a transition from a coherent-rotation behavior
to curling behavior, similar to Fig. 1. 11.

1. 5. 5 Cooperative Effects
Important aspects of nanomagnetism are magnetostatic and exchange
interaction between particles and grains. First, the determination of the
coupling constants from the real structure is a complicated problem, involving
for example effective RKKY interactions between grains (Skomski, 1999) and
an integration over grain-boundary spin structures such as that in Fig. 1. 16a
(Skomski et ai., 1998a, 2001). Second, the effect of these interactions on
the magnetic behavior of the nanostructures is often very important. For
example, in magnetic recording media they lead to the formation of interaction
domains, which may improve the thermal stability but reduce the storage
density. In permanent magnets the vanishing of the two-phase shoulders in
hysteresis loops can be considered as a cooperative effect (Liu et al. , 2000)
and the low coercivity of amorphous soft magnets is a cooperative random
anisotropy effect (Section 1. 5. 6) .
The simplest way to describe interactions between particles is to
introduce an interaction field. Examples of this mean-field type approach are
the Preisach model (Preisach, 1935) and various approaches (Henkel, 1964;
Veitch, 1990; Che and Bertram, 1992; Basso et ai., 1994) based on
Wohlfarth's remanence relation (Wohlfarth, 1958). However, interaction
fields are unable to give an appropriate description of cooperative
magnetization processes (Skomski and Sellmyer, 2001). Good fits of
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 41

experimental hysteresis loops do not necessarily indicate the absence of


cooperative effects, because even trivial equations such as
00

M(H) = M s - fMsP(H')dH' (1. 26)


H

can map any hysteresis loop onto a switching-field distribution P (H) .


Figure 1. 19 shows some cooperative nucleation modes for arrays of
particles with perpendicular anisotropy. The flux closure shown in (d) and (e)
is favorable from the point of view of magnetostatic energy and leads to a
nucleation-field reduction ignored in mean-field and Preisach-type approaches.
For the geometries Fig. 1. 19d and e the cooperative correction is comparable
to the uncorrected interaction field acting on the dots.

(a) (b) (c)


I


(d)
(e)

Figure 1.19 Cooperative effects in nanodot arrays: (a) degeneracy of the coherent-
rotation mode in a small particle. (b) breaking of the degeneracy by a defect, (c) two-
particle cooperative mode. (d) cooperative mode involving more than two dots, and
(e) flux closure as one origin of cooperative behavior.

A criterion for the applicability of Preisach-type models can be obtained


from the slope X ( He) of the hysteresis loop: when the interaction field
exceeds Msi X (He)' then the behavior of the magnet is governed by
cooperative effects (Skomski et al., 1998a, Skomski and Sellmyer, 2001).
Note that Msi X ( He) can be considered as the width of the switching-field
distribution. In terms of Fig 1. 19, this effect corresponds to the destruction of
the cooperative flux closure (d) and (e) due to single-particle disorder. For
example, imperfections such as that shown in (b) may lead to the low-field
switching of individual dots or grains.
42 R.Skomski and D. J. Sellmyer

1. 5. 6 Random Anisotropy and Remanence Enhancement


As mentioned in Section 1. 3. 1, isotropic ensembles of interaction-free small
particles can be considered as Stoner-Wohlfarth particles. However,
intergranular interactions lead to the breakdown of the picture of individual
grains. In particular, strong exchange coupl ing in isotropic magnets means the
macroscopic anisotropy averages to zero. On the other hand, the
ferromagnetic exchange J yields some alignment of the magnetization and an
increase in remanence known as remanence enhancement (Coehoorn et al. ,
1988; Manaf et al. , 1993).
On an atomic level, amorphous random-anisotropy is described by the
Harris-Plischke-Zuckerman model (Harris et al., 1973). The model and its
extensions lead to two types of hysteresis loops. In the "strong-pinning"
regime, where the atomic anisotropy K dominates the exchange, the atomic
spins behave non-cooperatively, the remanence is given by Mrl M s = 112 +
zJ/6K, and the coercivity approaches 0.479 H a (Section 1.3. 1). In the
"weak-pinning" regime, J K, there is a cooperative remanence
enhancement, Mr/M s = 1- (4/15) (KlzJ)2, and the very small coercivity
obeys a (KIJ)4power law (Callen et al., 1977; Alben et ai., 1978; Moorjani
and Coey, 1984). Since many amorphous materials are characterized by J
K, this random-anisotropy model is particularly relevant to soft magnets.
Random-anisotropy effects are of interest in nanomagnetism, because
interaction domains in amorphous magnets contain many atoms. On the other
hand, many nanostructured materials are isotropic, and their behavior exhibits
a characteristic grain-size and anisotropy dependence. Examples are isotropic
hard-soft nanostructures (Coehoorn et al., 1988; Manaf et al., 1993),
textured permanent magnets (Skomski et al., 1998a), patterned media
(Loffler et ai., 1999; Skomski et al., 2001), and polycrystalline nanowires
(Zheng et al., 2000; Skomski et al., 2000). The ground-state domain
structure of partially ordered nanomagnets is investigated in Chudnovsky et al.
( 1986), whereas the hysteretic behavior of these magnets has first been
discussed by Skomski et al. (1998a). In these theories, the strength of the
exchange is expressed in terms of the dimensionless parameter AI K 1 R 2 ,
where R is the average grain size, although grain-boundary effects
(Section 1.5.3) are important in the strong-coupling.
The strong-interaction Iimit is characterized by an interesting scal ing
behavior first investigated in the context of random-field magnetism (Imry and
Ma, 1975). Let us assume that the d-dimensional magnet forms domains of
size L d' each domain containing many randomly oriented crystallites of size
R. The exchange-energy density scales are AIL ~, whereas the anisotropy
energy is of the order K elf = K 1 (R 1L d) d-2. The net anisotropy K elf is much
smaller than K l ' because it reflects incomplete anisotropy averaging
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 43

(Gaussian fluctuations). Minimizing the total energy yields L d and K eff. The
domain size is given by
Ld """, 00(00/R)dl(4-dl (1.27)

where 00 = (A / K 1) 1/2. This means that domains become very large when R <
00' that is, for small grains and in the case of soft materials, where 00 is
large. On the other hand, tak ing into account that H cex:; K 1 , we find that
H c ""'" He (R / 00 )2dl(4-dl . (1.28)
This shows that the large domains are accompanied by very small coercivities.
This has been exploited in the development of nanocrystalline and amorphous
soft magnets (Herzer, 1995). Interestingly, Eqs. (1. 27) and (1. 28) reveal a
marginal dimension d = 4 above which the power laws degenerate into sharp
jumps.

1. 5. 7 Magnetization Dynamics in Nanostructures


The nonequilibrium character of magnetization processes means that
magnetization processes are time-dependent, even if the external magnetic
field is kept constant. Atomic processes lead to fast equilibrium on a local
scale and realize intrinsic properties on subnanosecond time scales. Intrinsic
properties are therefore equilibrium properties, and the energy functional
Eq. (1. 9) is also known as the micromagnetic free energy. However, extrinsic
magnetic properties are realized on length scales of several nanometers, and
thermal excitations are often unable to overcome the free-energy barriers
separating metastable states. For example, freshly magnetized permanent
magnets lose a small fraction of their magnetization within the first few hours.
Intermediate time scales play a role, for example, in magnetic resonance.
This section focuses on slow dynamics; fast atomic processes and phenomena
such as spin precession in domain walls and quantum-tunneling will be dealt
with elsewhere in this book.
The time-dependent many-body Schr6dinger equation ifia I IfI > at = H I
1fI> can, in principle, be used to predict the evolution of any physical
system, but this method is not feasible in practice. First, the deterministic
character of the Schr6dinger equation forbids irreversible processes. Second,
the large number of involved degrees of freedom, such as lattice vibrations,
complicates the description of real magnetic systems. To make meaningful
predictions about the relevant magnetic degrees of freedom, such as the
position of a domain wall, one must treat the irrelevant degrees of freedom as
a heat bath. This "coarse graining" (Zwanzig, 1961; Mori, 1965) has a
simple classical analogue. Consider a harmonic system of masses coupled by
springs. The system has a recurrence time Trecscaling as 1/ Dow, where Dow is
the system's smallest eigenfrequency difference. For any finite system the
44 R. Skomski and D. J. Sellmyer

recurrence time is finite. but for an infinite number of degrees of freedom.


corresponding to a heat bath. llw = 0 and T rec = 00 .
The coarsegraining procedure leads to or justifies various nonequilibrium
approximations. The Landau-Lifshitz equation
dM 1
-d
t
= YoM x H eff - ~M x (M x H eff ) (1.29 )
lVi s To
where To is an inverse attempt frequency of the order of 10 - 9 S (Neel. 1949;
Street and Wooley. 1949), and similar relations such as the Gilbert and Bloch-
Bloembergen equations contain both deterministic effects and relaxation.
Aside from the precession of the magnetization. they describe the viscous
rotation of M towards the effective field H eff = - 5E (M) 15( /Jo M ) , that is. the
relaxation in the vicinity of local or global energy minima. An equation
describing thermally activated jumps over energy barriers is the Langevin
equation:

~ = _ -..G.- aE + ,j2F;0_~(t) (1.30)


at k Ta5
s
where ~ is a magnetic phase-space vector and 10 = liTo is an atomic attempt
frequency. The random forces ~ (t) obey <~ ( t
= 0 and <~ (t) ~ ( > = n
5( t - t'). where 5 is the delta function. Note that Eq. ( 1. 30) disregards the
atomic spin precession, which is much faster than the macroscopic
equil ibration time.
The probability distribution belonging to Eq. (1. 30) obeys the diffusion-
type magnetic Fokker-Planck equation (Kramers, 1940; Brown, 1963b)

aP ( 5) = _1_ ~ (p aE ) + a2P (1.31)


To at ks T a~ a~ a 52
Both Eqs. (1. 30) and (1.31) can be derived from a phenomenological master
or rate equation

( 1. 32)

where the W(~.~') = W(~'-~) are appropriately chosen transition rates. In


equilibrium, Eqs. (1.30) - (1. 32) all reproduce the Boltzmann distribution
P(~)= (liZ) exp[- E (~)lksT]. A simplified version of the master
equation. valid for transitions between two energy minima. is a PIN t =
W 12 P 2 - W 21 PI and a P 21a t = - W 12 P 2 + W 21 PI' Kramers' escape-rate
theory (Kramers, 1940), originally used to describe chemical reaction
kinetics (Hanggi et al . 1990). derives the transition rates W 12 and W 21 ; they
are proportional to exp ( - E al k s T). In magnetism, Kramers' escape-rate
theory is also known as the Arrhenius-Neel theory.
The extrinsic time dependence of the magnetization is known as magnetic
viscosity. The magnetic viscosity determines, for example. the stability of the
information stored in magnetic and magneto-optic recording media (Sellmyer
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 45

et aI., 1998). It is caused by thermal activation and involves jumps over


(free) energy barriers. Relaxation in an ensemble of non interacting small
particles is characterized by the relaxation times (Becker and Doring, 1939;
Kneller, 1962)

T; = TOexp( - k~aT) (1.33)

where i is a particle index. The energy barriers E a (H) depend, for example,
on the shape and anisotropy of the particles (Becker and Doring, 1939;
Skomski and Christoph, 1989). In more complex systems, such as interacting
particles and bulk magnets, the index i refers to the eigenmodes of the
magnetization process (Skomski and Sellmyer, Victora, 1989). In the
simplest case, the magnetization of the particles is M; (t) = - M s + 2M s exp
(- t/ T;). For t =0 and t = 00 this equation yields M; = M s and M; = - M s '
respectively. The total magnetization M ( t) is obtained by averaging overall
M; ( t ). Introducing an energy-barrier distribution P ( E a ) , we have to
determine the integral

f
~

M(t) =- M s + 2M s P(Ea)e-WToJexP<-EalkBTJdEa' (1.34)

To realize the averaging we exploit the mathematical identities that aexp({3E) =


exp({3E+lna) andexp[-exp(-x!E)] =G(x), where El and G(x) is
the step function defined by G(x<O) =0 and G(x>O) = 1. The result is the
important logarithmic law
M(t) = M(t') - 2k s T P(O) lnCt/t')M s ( 1. 35)
where t and t' are two arbitrary times. Equations of this type were derived
many decades ago. An example is a calculation based on the unnecessary
assumption of a rectangular energy-barrier distribution of width W, so that
P(O)=l/W (Becker and Doring, 1939).
From a phenomenological point of view, the logarithmic law is often
written as M ( t) = M ( t) - 8 m In ( t/n,where 8 m is the magnetic-viscosity
constant (Becker and Doring, 1939; Kneller, 1962; Gaunt, 1986; Skomski
and Coey, 1999). Note that the logarithmic law is unphysical not only in the
short-time limit, where spin precession is important, but also for extremely
long times, where M z = - 00, rather than - Ms. The main reason is the
approximation P ( E) = 1/ Eo, which breaks down for very high energies
(Becker and Doring, 1939). A more adequate expression is (Skomski and
Christoph, 1989)

( 1. 36)
but since (x E - 1) = E Inx for small exponents, this expression is not very
different from the logarithmic law. For other non-logarithmic expressions see
(Aharoni, 1996) and (Berkov, 1992).
46 R. Skomski and D. J. Sellmyer

An important point is the dependence of the energy barriers on the


external field. There is a straightforward connection between the energy-
barrier density P, the irreversible part Xirr of the susceptibility, and field-
derivative of the energy barriers. Since dM / M s = 2 P dE, X irr = J M /CJ H, and
dE=(JE/JH) dH, the logarithmic law can also be written as (Street and
Wooley, 1949)

(1.37)

This equation is frequently used to rationalize energy-barrier effects (Street


and Wooley, 1949; Gaunt, 1986; Givord and Rossignol, 1996; Sellmyer et
al. , 1998).
An effect closely related to magnetic viscosity is the dependence of the
coercivity on the sweep rate dH / d t. He is largest for high sweep rates, that
is, for fast hysteresis-loop measurements (Sellmyer et al., 1998). Let us
assume that the energy barriers exhibit a power-law dependence

( 1. 38)

where m is discussed in Section 1. 5. 8 and K eff includes the particles' shape


anisotropy. The physical or "Barkhausen" volume V is not necessarily the
volume of a single particle. Due to cooperative and localization effects it may
be smaller or larger than the particle volume. Since T = To exp( E B / k B T) is
the time necessary to jump over the free-energy barrier, E B (He) = k B T In ( T/ To) ,
where T is the characteristic time scale of the coercivity measurement. This
yields (Kneller, 1966; Sharrock, 1994)

He = H A {l- [:e:~lnCT/To)rm}. (1.39)

Improving on Eq. ( 1. 39) by a master-equation approach yields a relatively


unimportant factor of In 2 =O. 693 (Kneller, 1966), which is usually
incorporated into TO' In practice, T is tuned by varying the sweep rate 1') =
dH / d t - 1/ T. An experimental approach to analyze the resulting sweep-rate
dependence of the coercivity is to exploit the phenomenological relation
(Sellmyer et al. , 1998)

= H eO
(1') ) + k B T
MsV'
In (!l)
1')0
( 1. 40)

where V' is the activation volume. This equation amounts to a fluctuation-field


or sweep-rate correction (Neel, 1951) to the static coercivity H eO; it is
essentially equivalent to the !:>.H term in Eq. (1. 17). Linearizing (1. 39) with
respect to In( 1') /1')0) = - In( T/ To) and comparing the result with ( 1. 40) yields
(Skomski and Coey, 1999)

(1.41>
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 47

This equation shows that the activation volume V' is only loosely related to
the" physical" volume V. Note that V' contains two types of temperature
dependences: an intrinsic temperature dependence via K 1 (T) (Skomski and
Coey, 1999), and an extrinsic temperature dependence via Tl/ m .
An alternative method to derive Eq. ( 1. 41) is to start from the expression
(Street and Wooley, 1949; Givord et ai., 1987, 1990; Sellmyer et al. ,
1998, 1999)

V' - 1 aEB(H) I ( 1. 42)


Ms aH He(T.V)

where He is obtained by solving E B (He) = 25 k B T and E B is given by


Eq.(1.38). For m =2, the activation volume V' = V(l- HeIH A ). The
relation between V' and S is given by V' = k B T 1M s SXirr (Street and
Wooley, 1949; Givord et ai., 1990; Sellmyer et ai., 1998, 1999). Both S
and Xirrgenerally have peaks near He' so that V' (He) is often determined
only at the single field value where the reversal is most significant. S values
determined from magnetic-viscosity experiments tend to agree reasonably well
with sweep-rate values of S.
Equation (1. 38) shows that free-energy barriers decrease with the
particle size. When K 1 V is comparable to k B T the magnetization decays very
rapidly, which is known as superparamagnetism. Defining superparamagnetism by
T = 100 s leads to the stability condition K 1 VI k B T~25, where K 1 VI k B T= ~ is

referred to as the stability parameter. The magnetization of a grain of volume V =


(5 nm)3 and anisotropy K u = 4 X 106 erg/cm3 is thermally unstable at room
temperature, because ~ = 12. On the other hand, for a grain to be stable for
10 years (3 x lOB s), ~~40. For particles whose radius is smaller than about
1 nm the external magnetic field is unable to produce saturation, because it
cannot compete against thermal excitations. Note that true ferromagnetism is
limited to infinite magnets, because thermal excitations in finite magnets cause
the net moment to fluctuate between opposite directions and yield, ultimately,
a zero spontaneous magnetization. However, due to the long equilibration
times it is difficult to distinguish the magnetism of small particles from true
ferromagnetism.

1. 5. 8 Energy-Barrier Laws

A complicated problem is the relation between energy barriers such as


Eq. (1. 38) and the real structure of a magnet. For aligned and non-interacting
Stoner-Wohlfarth grains, it is straightforward to show that Eq. (1. 38) is
correct with m = 2 and K eff = K 1 More generally, the exponent m varies
between 3/2 to 2 for a variety of pinning and nucleation models (Gaunt, 1986;
Victora, 1989). The exponent m = 3/2, which was first derived by Neel
(1950), is quite common and describes a variety of coherent and incoherent
48 R. Skomski and D. J. Sellmyer

magnetization processes. Examples are strong domain-wall pinning, as in


Fig. 1. 12, the reversal of misaligned Stoner-Wohlfarth particles (Victora,
1989; Gaunt, 1983).
Linear laws, where m = 1, are very popular (Gaunt, 1983) but have no
sound physical basis. Nonanalytic energy landscapes E (x) could, in
principle, yield m = 1 (Gaunt, 1983), but so far it hasn't been possible to
derive them from realistic energy landscapes. For example, sharp interfaces
and other atomically localized features do not establish nonanalytic
singularities, because the barrier energy E (x) then reflects the smooth
domain-wall fine structure and is proportional to 1- tanh z (x / 00)' where 00 =
(A / K , ) I/Z. Other approaches start from unreal istic (Egami, 1973a, 1973b)
or ill-defined energy landscapes and yield unreasonable predictions, such as
an infinite zero-temperature coercivity.
A simple derivation, based on ideas developed in catastrophe theory,
(Pinto, 1987), is to use an expansion of the micromagnetic (nanomagnetic)
energy, as outl ined in Fig. 1. 8a. Including linear, quadratic and cubic terms,
the energy can be written as

E(x) = 00 + o,x + ~zxz + ~3X3 - boHx (1.43)

where x describes the degree of freedom associated with the relevant reversal
mode. Stoner-Wohlfarth particles are characterized by the coherent-rotation
angle x = e, whereas in the case of strong domain-wall pinning x describes
the wall-position (Fig. 1. 12). The phenomenological parameters 00' 0" Oz,
03 and b o describe the real-structure of the magnet and must be determined
separately.
Analysis of Eq. (1. 43) yields a local minimum vanishing at the static
switching field H co = 01/ b o - oU 403 b o and reproduces Eq. ( 1. 38) with m =
3/2. It is important to keep in mind that Eq. (1. 43) includes localization
(Victora, 1989), interaction (Lottis et al., 1991) and cooperative (Skomski
and Sellmyer, 2001) effects, although the specification of these mechanisms
can be very difficult. Deviations from Eq. ( 1. 43) occur for highly symmetric
energy landscapes, where 03 = O. It is then necessary to include fourth-order
terms, and the power-law exponent changes to m = 2. A typical example is
aligned Stoner-Wohlfarth particles, whose energy Eq. (1. 11) does not contain
odd-order terms.
As an example, we will consider the localized reversal in nanowires. In
addition to the defect t:.K discussed in Section 1. 5. 4, we add some easy-axis
misalignment e (x) = Leo 0 (x). Figure 1.20 illustrates the meaning of this
defect. Evaluating the magnetic energy with the help of the trial function (x) = 0 .
exp(-j x I/R L ) yields an energy expression similar to Eq. (1.43) (see
Fig.1.20) m=3/2, and the coercivity (Zeng et al., 2002)

H = 2KeM[1_~_3(2Leo/RL)Z/3J (1.44)
c /JoM s KR~ 4
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 49

This equation shows that both the reduced local anisotropy f!.K (via L R) and
the grain misalignment (eo) contribute to the experimentally observed
coercivity reduction, although the present model does not distinguish between
shape, bulk, and surface contributions to f!.K. The physical activation
volume, V= 2nR 2 R L , is a function of A. For eo = 0, the procedure yields
m=2.

::i
~
l<i

I
!
t
(a) (b)

Figure 1. 20 Localized reversal in a nanowire: Ca) geometry and Cb) energy landscape.

1. 6 Conclusions

In this introductory chapter, we have discussed fundamental aspects of


advanced magnetic materials. The important distinction between intrinsic and
extrinsic properties has been elaborated, and it has been discussed how this
distinction is realized in magnetic nanostructures. Intrinsic properties, such as
spontaneous magnetization, Curie temperature and magnetocrystall ine
anisotropy, are realized on atomic length scales and therefore well-defined for
nanostructures. On the other hand, extrinsic properties, such as coercivity,
are real ized on larger length scales and strongly real-structure dependent.
A widely encountered nanoscale extrinsic effect is micromagnetic
local ization. It is caused by imperfections, occurs in a broad variety of
materials, and affects static and dynamic magnetization processes. An explicit
analysis has been provided for bulk composites and nanowires. From the
quantum-mechanical analogy it follows that an arbitrarily weak inhomogenity
leads to the localization of all magnetic modes in one-dimensional magnets
(thin nanowires), but for zero disorder the localization length goes to infinity
and the reversal degenerates into coherent rotation.
50 R. Skomski and D. J. Sellmyer

This chapter has shown how atomic-scale physics can be exploited to


create magnetic materials with well-defined properties and how these
properties are used in magnetic applications. The search for materials with
improved intrinsic properties continues to be of scientific and technological
interest, but the main thrust of magnetic materials is the creation,
understanding, and exploitation of artificial nanostructures. In the near future,
this materials-by-design concept will be of key importance in the science and
technology of magnetic materials.

Appendix 1. 1 Magnetic Units

In a strict sense, the 81 unit for the magnetization M and for the magnetic field
H is Aim. This reflects the 19th-century belief that the magnetization is caused
by electric currents. (The magnetization of transition-metal magnets is largely
due to the spin of the electron and can therefore not be reduced to currents
characterized by v < c.) The magnetic flux density, or induction, B and the
flux-density equivalents /-10 M and /-10 H are measured in tesla (T). The quantity
J = /-10 M is also known as the magnetic polarization, but this J is easily
confused with exchange constants and angular momenta. Occasionally, the
field is denoted by B a or B o , but this blurs the physical important distinction
between the magnetic field and the induction and leads to cumbersome
subscripts for derived quantities such as energy products and interaction
fields.
One advantage of the unit of tesla is its easy conversion into cgs units:
one tesla is equal to 10 kG (flux density 4n M) and to 10 kOe (field H). The
cgs magnetization unit is emul cm 3 (1 emul cm 3 = 1 kA/m) , so that the
multiplication by the dimensionless constant 4n leads to a cgs unit change,
from emu/cm3 (M) to G (4nM). Other useful conversions are 1 Oe=79. 577 Aim
(field), 1 emu/g= 1 J/Tkg (mass polarization), and 1 MGOe= 7.9577 kJ/m 3
(energy product). Finally, it is useful to keep in mind the 81 relations
B = /-10 H + /-10 M, E z = - /-10 MH, and 0 = 1 correspond to the cgs relations
B = H + 4nM, E z = - MH, and N = 4n.

References
Aharoni, A. Rev. Mod. Phys. 34: 227 (1962)
Aharoni, A. Introduction to the Theory of Ferromagnetism. University Press,
Oxford (1996)
Alben, R., J. J. Becker and M. C. Chi. J. Appl. Phys. 49: 1653 (1978)
Anderson, P. VV. Phys. Rev. B 109: 1492 (1958)
Ashcroft, N. VV. and N. D. Mermin. Solid State Physics. Holt, New York
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 51

(1976)
Baibich, M. N. , J. M. Broto, A. Fert, F. Nguyen van Dau, F. Petroff, P.
Etienne, G. Creuzet, A. Friederich and J. Chazelas. Phys. Rev. Lett. 61:
2472 (1988)
Bander, M. and D. L. Mills. Phys. Rev. B 38: 12,015 (1988)
Basso, V., M. LoBue and G. Bertotti. J. Appl. Phys. 75: 5677 (1994)
Becker, R. and W. Doring. Ferromagnetismus. Springer, Berlin (1939)
Berkov, D. V. J. Magn. Magn. Mater. 111: 327 (1992)
Berkowitz, A. E., J. R. Mitchell, M. J. Carey, A. P. Young, S. Zhang, F.
E. Spada, F. T. Parker, A. Hutten and G. Thomas. Phys. Rev. Lett. 68:
3745 (1992)
Binder, K. Eur. J. Phys. 20: 389 (1999)
Bland, J. A. C. and B. Heinrich (eds.). Ultrathin Magnetic Structures I.
Springer, Berl in (1994)
Bloch, F. Z. Phys. 57: 545 (1929)
Bloch, F. Z. Phys. 61: 206 (1930)
Bloch, F. Z. Phys. 74: 295 (1932)
Bloch, F. andG. Gentile. Z. Phys. 70: 395 (1931)
Bowden, G. J., J. M. L. Beaujour, S. Gordeev, P. A. J. de Groot, B. D.
Rainford and M. Sawicki. J. Phys. : Condens. Matter 12: 9335 (2000).
Bozorth, R. M. Ferromagnetism. van Nostrand, Princeton (1951)
Braun, H.-B. and N. H. Bertram. J. Appl. Phys. 75: 4609 (1994)
Brooks H. Phys. Rev. 58: 909 (1940)
Brown W. F. Rev. Mod. Phys. 17: 15 (1945)
Brown W. F. Micromagnetics. Wiley, New York (1963a)
Brown W. F. Phys. Rev. 130: 1677 (1963b)
Bruno P. Phys. Rev. B 39: 865 (1989)
Brush S. G. Rev. Mod. Phys. 39: 883 (1967)
Callen E., Y. J. Liu and J. R. Cullen. Phys. Rev. B 16: 263 (1977)
Charles S. W. In: J. L. Dormann and D. Fiorani, eds. Studies of Magnetic
Properties of Fine Particles and their Relevance to Materials Science.
Elsevier, Amsterdam, p.267 (1992)
Che X.-D. and N. H. Bertram. J. Magn. Magn. Mater. 116: 121 (1992)
Chikazumi S. Physics of Magnetism. Wiley, New York (1964)
ChouS. Y., M. Wei, P.R. KraussandP. B. Fisher. J. Vac. Sci. TechnolB
12: 3695 (1994)
Chuang D. S., C. A. Ballentine and R. C. Q'Handley. Phys. Rev. B 49:
15,084 (1994)
Chudnovsky E. M. , W. M. Saslow and R. A. Serota. Phys. Rev. B 33: 251
(1986)
Coehoorn R., D. B. de Mooij, J. P. W. B. Duchateau and K. H. J.
Buschow. J. de Physique 49, C-8669 (1988)
Coehoorn R. Phys. Rev. B 39: 13,072 (1989)
Coey J. M.D. Physica Scripta T39: 21 (1991)
52 R.Skomski and D. J. Sellmyer

Coey J. M. D. Rare-earth Iron Permanent Magnets. University Press, Oxford


(1996)
Coey J. M. D. and R Skomski. Physica Scripta 1'49: 315 (1993)
Coey J. M. D. and H. Sun. J. Magn. Magn. Mater. 87: L251 (1990)
Comstock R. L. Introduction to Magnetism and Magnetic Recording. Wiley,
New York (1999)
Craik D. J. and R. S. Tebble. Rep. Prog. Phys. 24: 116 (1961>
Craik D. J. Magnetism: Principles and Applications. Wiley, New York
(1995)
Croat J. J. , J. F. Herbst, R. W. Lee and F. E. Pinkerton. J. Appl. Phys.
55: 2078 (1984)
Cyrot-Lackmann F. J. Phys. Chem. Solids 29: 1235 (1968)
Daalderop G. H. O. , P. J. Kelly and M. F. H. Schuurmans. Phys. Rev. B
42: 7270 (1990)
Desjonqueres M. C. and D. Spanjaard. Concepts in Surface Physics.
Springer, Berlin (1993)
Ding J., P. G. McCormick and R. Street, J. Magn. Magn. Mater. 124: L1
( 1993)
Dormann J. L. , D. Fiorani and E. Trone. In: G. C. Hadjipanayis and R. W.
Sigel eds. Nanophase Materials. Kluwer, Dordrecht, p.365 (1994)
Egami T. Phys. Stat. Sol. (a) 19: 747 (1973a)
Egami T., Phys. Stat. Sol. (a) 20: 157 (1973b)
Evetts J. E. Concise Encyclopedia of Magnetic and Superconducting
Materials. Pergamon, Oxford (1992)
Fert A. and L. Piraux, J. Magn. Magn. Mater. 200: 338 (1999)
Fischer R. and H. Kronmuller. J. Magn. Magn. Mater. 184: 166 (1998)
Fukunaga H. and H. Inoue. Jpn. J. Appl. Phys. 31: 1347 (1992)
Fullerton E. E. , S. J. Jiang and S. D. Bader, J. Magn. Magn. Mater. 200:
392 (1999)
Gaunt P. Phil. Mag. B 48: 261 (1983)
Gaunt P. J. Appl. Phys. 59: 4129 (1986)
Gay J. G. and R. Richter. Phys. Rev. Lett. 56: 2728 (1986)
Givord D. , A. Lienard, P. Tenaud and T. Viadieu. J. Magn. Magn. Mater
67: L281 (1987)
Givord D. , Q. Lu, M. F. Rossignol, P. Tenaud and T. Viadieu. J. Magn.
Magn. Mater. 83: 183 (1990)
Givord D. and M. F. Rossignol., J. M. D. Coey ed., Rare-Earth Iron
Permanent Magnets. University Press, Oxford. (1996)
Gradmann U. In: K. H. J. Buschow, ed. Handbook of Magnetic Materials,
Vol. 7. Elsevier, Amsterdam, p. 1 (1993)
Hadjipanayis G. C. J. Magn. Magn. Mater. 200: 373 (1999)
Hanggi P., P. Talkner and M. Borkovec. Rev. Mod. Phys. 62: 251 (1990)
Harris R., M. Plischke and M. J. Zuckermann. Phys. Rev. Lett. 31: 160
(1973)
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 53

Hehn M., K. Ounadjela, J. Bucher, F. Rousseaux, D. Decanini, B.


Bartenlian and C. Chappert, Science 272: 1782 (1995)
Heine V. Sol id State Phys. 35: 1 (1980)
Heisenberg W. Z. Phys. 49: 619 (1928)
Henkel D. Phys. Stat. Sol. 7: 919 (1964)
Herbst J. F. Rev. Mod. Phys. 63: 819 (1991)
Herzer G. Scripta Metal. 33: 1741 (1995)
Hertel R. J. Appl. Phys. 90: 5752 (2001)
Hilzinger H. R. and H. Kronmuller. Phys. Lett. A 51: 59 (1975)
Himpsel F. J. , J. E. Ortega, G. J. Mankey and R. F. Willis. Adv. Phys.
47: 511 (1998)
Hutchings M. T. Solid State Phys. 16: 227 (1964)
Huysmans G. T. A. and J. C. Lodder. J. Appl. Phys. 64: 2016 (1988)
Imry Y. and S.-K. Ma. Phys. Rev. Lett. 35: 1399 (1975)
Ising E. Z. Phys. 31: 253 (1925)
Jessensky 0., F. Muller and U. Gosele. J. Electrochem. Soc. 145: 3735
(1998)
Johnson M. T. , P. J. H. Bloemen, F. J. A. den Broeder and J. J. de Vries.
Rep. Prog. Phys. 59: 1409 (1996)
Kawai S. and R. Ueda. J. Electrochem. Soc. 122: 32 (1975)
Kersten M. Z. Phys. 44: 63 (1943)
Kittel C. Rev. Mod. Phys. 21: 541 (1949)
Kneller E. Ferromagnetismus. Springer, Berlin (1962)
Kneller E. In: H. P. J. Wijn, ed. Handbuch der Physik XIII/2:
Ferromagnetismus. Springer, Berlin, p.438 (1966)
Kneller E. and R. Hawig. IEEE Trans. Magn. 27: 3588 (1991)
Kramers H. A. Physica 7: 284 (1940)
KronmQller H. phys. stat. sol. (b) 144: 385 (1987)
Kronmuller H., K.-D. Durst and M. Sagawa. J. Magn. Magn. Mater. 74:
291 (1988)
Kuhrt C., K. 0' Donnell, M. Katter, J. Wecker, K. Schnitzke and L.
Schultz. Appl. Phys. Lett. 60: 3316 (1992)
Kumar K. J. Appl. Phys. 63: R13 (1988)
Landau L. and E. Lifshitz. Phys. Z. Sowjetunion 8: 153 (1935)
Lesl ie-Pelecky D. L. and R. L. Schalek. Phys. Rev. B 59: 457 (1999)
Li F.-Y. and R. M. Metzger. J. Appl. Phys. 81: 3806 (1997)
Liu J. P., C. P. Luo, Y.Liu and D. J. Sellmyer. Appl. Phys. Lett. 72: 483
(1998)
Liu J. P., R. Skomski, Y. Liu and D. J. Sellmyer. J. Appl. Phys. 87: 6740
(2000)
Loffler J. , H. -B. Braun and W. Wagner. J. Appl. Phys. 85: 5187 (1999)
Long G. J. and F. Grandjean. Supermagnets. Hard Magnetic Materials.
K luwer, Dordrecht (1991)
Lottis D. K. , R. M. White and E. D. Dahlberg. Phys. Rev. Lett. 67: 2362
54 R.Skomski and D. J. Sellmyer

( 1991)
Lyberatos A. J. Phys. D: Appl. Phys. 33: Rl17-R133 (2000)
Maciejewski W. and A. Duda. Solid State Commun. 64: 557 (1987)
Manaf A. , P. A. Buckley and H. A. Davies. J. Magn. Magn. Mater. 128:
302 (1993)
Martin C. R. Science 266: 1961 (1994)
Masuda H. and K. Fukuda. Science 268: 1466 (1995)
McCurrie R. A. Ferromagnetic Materials - Structure and Properties.
Academic Press, London (1994)
Millev Y., R. Skomski and J. Kirschner. Phys. Rev. B 58: 6305 (1998)
Mizuki I., Y. Yamamoto, T. Yoshino and N. Baba. J Metal Surf. Finish.
Soc. Jpn. 38: 561 (1987)
Moorjani K. and J. M. D. Coey. Magnetic Glasses. Elsevier, Amsterdam
(1984)
Mori H. Prog. Theor. Phys. 33: 423 (1965)
Neel L. Ann. Geophys. 5: 99 -136 (1949)
Neel L. J. de Phys. Rad. 11: 49 (1950)
Neel L. J. de Phys. Rad. 12: 339 (1951)
Neel L. J. de Phys. Rad. 15: 376 (1954)
Nieber S. and H. Kronmuller. Phys. Stat. Sol. (b) 153: 367 (1989)
Nielsch K. , R. B. Wehrspohn, J. Barthel, J. Kirschner, U. G6sele, S. F.
Fischer and H. Kronmuller. Appl. Phys. Lett. 79: 1360 (2001)
Oepen H. P. and J. Kirschner. Phys. Rev. Lett. 62: 819 (1989)
Ozin G. A. Adv. Mat. 4: 612 (1992)
Parhofer S. M., J. Wecker, Ch. Kuhrt, G. Gieres and L. Schultz. IEEE
Trans. Magn. 32: 4437 (1996)
Pinto M. A. Phys. Rev. Lett. 59: 2798 (1987)
Preisach F. Z. Phys. 94: 277 (1935)
Ross C. A. , R. Chantrell, M. Hwang, M. Farhoud, T. A. Savas, Y. Hao,
H. I. Smith, F. M. Ross, M. Redjdal and F. B. Humphrey. Phys. Rev. B
62: 14,252 (2000)
Routkevitch D. , A. A. Tager, J. Haruyama, D. Almawlawi, M. Moskovits
and J. M. Xu. IEEE Trans. Magn. 43: 1646 (1996)
Roxlo B., H. W. Deckman, J. Gland, S. D. Cameron and R. Cianelli.
Science 235: 1629 (1987)
Sabiryanov R. F. and S. S. Jaswal. J. Magn. Magn. Mater. 177-181: 989
(1998a)
Sabiryanov R. F. and S. S. Jaswal. Phys. Rev. B 58: 12,071 (1998b)
Sagawa M., S. Fujimura, N. Togawa, H. Yamamoto and Y. Matsuira. J.
Appl. Phys. 55: 2083 (1984)
Sagawa M., S. Hirosawa, H. Yamamoto, S. Fujimura and Y. Matsuura.
Jpn. J. Appl. Phys. 26: 785 (1987)
Saito M., M. Kirihara, T. Taniguchi and M. Miyagi. Appl. Phys. Lett. 55:
607 (1989)
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 55

Sander D., R. Skomski, C. Schmidthals, A. Enders and J. Kirschner. Phys.


Rev. Lett. 77: 2566 (1996)
Sandratskii L. M. Adv. Phys. 47: 91 (1998)
Sawicki M. , G. J. Bowden, P. A. J. de Groot, B. D. Rainford, J. M. L.
Beaujour, R. C. C. Ward and M. R. Wells. Phys. Rev. B 62: 5817
(2000)
Schrefl T. and J. Fidler. J. Magn. Magn. Mater. 177-181: 970 (1998)
Schrefl T. , J. Fidler and H. KronmOller. Phys. Rev. B 49: 6100 (1994)
Sellmyer D. J., M. Yu, R. A. Thomas, Y. Liu and R. D. Kirby. Phys. Low-
Dim. Struct. 1-2: 155 (1998)
Sellmyer D. J., M. Yu and R. D. Kirby. Nanostructured Mater. 12: 1021
(1999)
Sellmyer D. J., C. P. Luo, Y. Qiang and J. P. Liu. In: H. S. Nalwa, ed.
Handbook of Thin Film Materials, vol. 5: Nanomaterials and Magnetic
Thin Films. Academic Press, San Diego, p.337 (2002)
Sellmyer D. J. , M. Zheng and R. Skomski. J. Phys.: Condens. Matter 13:
433 (2001)
Sharrock M. P. J. Appl. Phys. 76: 6413 (1994)
Shen J., R. Skomski, M. Klaua, H. Jenniches, S. S. Manoharan and J.
Kirschner. Phys. Rev. B 56: 2340 (1997)
Shiraki M., Y. Wakui, T. Tokushima and N. Tsuya. IEEE Trans Magn. 21:
1465 (1985)
Skomski R. and V. Christoph. Phys. Stat. Sol. (b) 156: K149 (1989)
Skomski R. Phys. Stat. Sol. (b) 174: K77 (1992)
Skomski R. and J. M. D. Coey. Phys. Rev. B 48: 15812 (1993)
Skomski R. J. Appl. Phys. 83: 6503 (1998)
Skomski R. , J. P. Liu, J. M. Meldrim and D. J. Sellmyer. In: L. Schultz
and K. -H. MOiler, eds Magnetic Anisotropy and Coercivity in Rare-Earth
Transition Metal Alloys. Werkstoffinformationsgesellschaft, Frankfurt/M.,
p.277 (1998b)
Skomski R., H.-P. Oepen and J. Kirschner. Phys. Rev. B 58: 11,138
(1998c)
Skomski R. , H. -P. Oepen and J. Kirschner. Phys. Rev. B 58: 3223 (1998d)
Skomski R. , C. Waldfried and P. A. Dowben. J. Phys.: Condens. Matter
10: 5833 (1998e)
Skomski R. Europhys. Lett. 48: 455 (1999)
Skomski R. and J. M. D. Coey. Permanent Magnetism. Institute of Physics,
Bristol (1999)
Skomski R. , J. P. Liu and D. J. Sellmyer. Phys. Rev. B 60: 7359 (1999)
Skomski R. and D. J. Sellmyer. J. Appl. Phys. 87: 4756 (2000)
Skomski R., H. Zeng, M. Zheng and D. J. Sellmyer. Phys. Rev. B 62:
3900 (2000)
Skomsk i R., In: M. Ziese and M. J. Thornton, eds. Spin-Electronics.
Springer, Berlin, p.204 (2001)
56 R. Skomski and D. J. Sellmyer

Skomski R. and D. J. Sellmyer. J. Appl. Phys. 89: 7263 (2001)


Skomski R., H. Zeng and D. J. Sellmyer. IEEE Trans. Magn. 37: 2549
(2001)
Slater J. C., Rev. Mod. Phys. 25: 199 (1953)
Smart J. S. Effective Field Theories of Magnetism. Sounders, Philadephia
(1966)
Steinbeck L., M. Richter, U. Nitzsche and H. Eschrig. Phys. Rev. B 53:
7111 (1996)
Stoner E. C. and E. P. Wohlfarth. Phil. Trans. Roy. Soc. A240: 599 (1948)
Street R. and J. C. Wooley. Proc. Phys. Soc. A62: 562 (1949)
Sugawara A. , D. Streblechenko, M. McCartney and M. R. Scheinfein. IEEE
Trans. Magn. 34: 108l( 1998)
Sutton A. P. Electronic Structure of Materials. Oxford University Press,
Oxford (1993)
Veitch R. J. IEEE Trans. Magn. 26: 1876 (1990)
Velu E. M. T. and D. N. Lambeth. J. Appl. Phys. 69: 5175 (1991)
Versluijs J. J., M. A. Bari and J. M. D. Coey. Phys. Rev. Lett. 87: 026,
601, 1 (2001)
Victora R. H. Phys. Rev. Lett. 63: 457 (1989)
VictoraR. H. andJ. M. McLaren. Phys. Rev. B47: 11,583 (1993)
Wang D.-S., R.-Q. Wu and A. J. Freeman. Phys. Rev. B 47: 14,932
(1993)
Warin P. , R. Hyndman, J. Glerak, J. N. Chapman, J. Ferre, J. P. Jamet,
V. Mathet and C. Chappert. J. Appl. Phys. 90: 3850 (2001)
Weller D. , A. Moser, L. Folks, M. E. Best, W. Lee, M. F. Toney, M.
Schwickert, J. -U. Thiele and M. F. Doerner. IEEE Trans. Magn. 36: 10
(2000)
Wernsdorfer W., K. Hasselbach, A. Benoit, B. Barbara, B. Doudin, J.
Meier, J.-P. Ansermet and D. Mailly. Phys. Rev. B 55: 11,552 (1997)
Wirth S., M. Field, D. D. Awschalom and S. von Molnar. Phys. Rev. B 57:
R14,028 (1998)
Wohlfarth P. J. Appl. Phys. 29: 595 (1958)
Wood R. IEEE Trans. Magn. 36: 36 (2000)
Xiao J. Q. , J. S. Jiang and C. L. Chien. Phys. Rev. Lett. 68: 3749 (1992)
Yeomans J. M. Statistical Mechanics of Phase Transitions. University Press,
Oxford (1992)
Zangari G. and D. N. Lambeth. IEEE Trans. Magn. 33: 3010 (1997)
Zeng H. , M. Zheng, R. Skomski, D. J. Sellmyer, Y. Liu, L. Menon and S.
Bandyopadhyay. J. Appl. Phys. 87: 4718 (2000)
Zeng H., R. Skomsk i, L. Menon, Y. Liu, S. Bandyopadhyay and D. J.
Sellmyer. Phys. Rev. B 65: 134,426 (2002)
Zheng M., R. Skomsk i, Y. Liu and D. J. Sellmyer. J. Phys.: Condens.
Matter. 12: L497 (2000)
Zheng M. , M. Yu, Y. Liu, R. Skomski, S. H. Liou, D. J. Sellmyer, V. N.
Intrinsic and Extrinsic Properties of Advanced Magnetic Materials 57

Petryakov, Y. K. Verevkin, N. I. Polushkin and N. N. Salashchenko.


Appl. Phys. Lett. 79: 2606 (2001)
Zhou J. , R. Skomski, C. Chen, G. C. Hadjipanayis and D. J. Sellmyer.
App!. Phys. Lett. 77: 1514 (2000)
Ziese M. and M. J. Thornton (Eds.). Spin-Electronics. Springer, Berlin
(2001)
Zwanzig R. Phys. Rev. 124: 983 (1961).

The authors wish to thank P. A. Dowben, R. D. Kirby, D. Leslie-Pelecky, S. -H. Liou, J.


P. Liu and J. Zhou for useful discussions of scientific topics involved. This work has been
supported by DOE, ONR, DARPA, AFOSR, NIR and CMRA.
2 Magnetism in Ultrathin Films and Beyond

Dongqi Li

2. 1 Introduction

The scheme of modern magnetism is spin-engineered nanostructures. The


demands in understanding miniaturization of magnetic systems directly stem
from the magnetic recording industry, where the bit size is already< 100 nm
and is continuously shrinking at a fast rate as information storage density
increases annually. The emerging spintronics, which employ the spin
component in electronics, further utilize magnetic nanostructures and may lead
to a revolution in microelectronics (Hathaway and Prinz, 1995; Prinz, 1999).
Meanwhile, the small magnetic structures also impose fundamental scientific
challenges by introducing exciting new phenomena. The physical size of a
magnetic system affects its magnetic properties by altering its dimensionality,
structure, surface/interface electronic structure, quantum size effects and
transport, domain wall structure and motion, etc. Some examples of size-
dependent physical phenomena and their characteristic lengths are listed in
Fig.2. 1. With the maturing of epitaxial film growth techniques, such magnetic
phenomena in two-dimensional systems are being extensively investigated with
thin films and multi layers . In addition, a new forefront is emerging in
fabricating laterally confined films via self-assembly and characterizing their
new physical properties. As a fast evolving field, magnetic ultrathin films and
nanostructures have been the subject of numerous reviews (for example,
Himpsel et ai., 1998; Sanders et al., 1999). In this chapter, we will
emphasize the most recent developments. First, fabrication of magnetic
ultrathin films and the related lateral structures such as wires and dots will be
discussed. Then we will use several recent examples to illustrate how
magnetic properties are strongly altered by the nanoscale sizes of the
systems. Specifically, we will mainly focus on the metastable phases of face-
centered Fe on Cu (100), finite size scaling in step-decorated Fe stripes on
vicinal Pd (110), surface electronic structure and magnetic properties of Gd
( 000 1 ), spin-polarized quantum well states in giant magnetoresistance
multi layers , and self-assembled magnetic dots, antidots, dot chains, and
stripes in epitaxial Co on Ru (000 1). The aim is to understand the general
research trends in this field and to look forward to a bright future of nanoscale
magnetic materials by design.
Magnetism in Ultrathin Films and Beyond 59

Figure 2. 1 Illustration of different size-dependent magnetic phenomena and the corresponding


characteristic lengths.

2. 2 Fabrication

The advents of nanomagnetism field critically depend on the progress in state-


of-art fabrication capabilities for well-defined small magnetic structures.
During the past one or two decades, the science and art of epitaxial growth of
magnetic thin films has been in the process of being perfected with molecular
beam epitaxy (MBE) and, to a lesser extent, sputtering (Fullerton et al. ,
1996). Most recently, there has been a strong interest in fabricating lateral
magnetic structures, either with Iithography or self-assembly. Lithography has
become a subfield of its own, and many exotic self-assembly mechanisms are
being explored. In this chapter, we will review the fundamentals of film
growth. Then we will extend these principles to growth-based self-assembly of
magnetic wires and dots, which normally have good crystalline ordering with
low defect density both in bulk and at boundaries.

2. 2. 1 Ultrathin Films

Good film growth is characterized by smooth layer-by-layer growth, a high


degree of epitaxial ordering and sharp interface. Near equilibrium, typically
real ized at relatively high temperature and/or slow growth rate, the
morphology of a film is determined by the surface free energy y of the
substrate, the film and their interface energy (Bauer and Kristallogr, 1958).
60 Dongqi Li

Thin crystalline films can therefore grow in one of the three modes: the Vomer-
Weber (VW) or 3D growth, the Frank-van der Merwe (FM) mode (or layer-
by-layer mode) and the Stranski-Krastanov (SK) mode as illustrated in Fig.
2.2. Layer-by-Iayer growth (FM mode) occurs only when !:ly= Yf+ Yi- Ys
~O for all thickness, where Yf' Ys and Yi are the surface energies of the
film, the substrate and their interface energy, and there is no strain from
lattice mismatch in the film. Experimentally, layer-by-layer growth mode with
monolayer islands is normally characterized with intensity oscillations of
reflected/ diffracted electron, atom, ion or X-ray beams, especially with
reflecting high-energy electron diffraction (RHEED). Each period indicates the
completion of a monolayer. The existence of misfit strain results in either high
density of dislocations or islands following the initial wetting layers, which will
be further discussed in fabricating dots. For conventional epitaxial growth,
lattice match between overlayer and substrate is crucial in producing high-
quality, smooth, well-ordered films.

2D S-K 3D

Figure 2.2 Schematics of the three equalibrium growth modes.

When smooth growth is not favored in equilibrium, non-equilibrium growth


can be employed. For example, one can grow films at a low temperature or
high rate to keep the films flat, and then anneal them to reduce the density of
structural defects. Another technique is to use a surfactant, a low surface
energy material that floats to the top of a film during growth (Egelhoff and
Jacob, 1989; Wuttig et al., 1993). For example, as seen in Fig.2. 3, clean
fcc Fe initially grow onto Cu (100) smoothly for 10 - 12 mono layers before
turning into island growth of bcc Fe. With certain surfactants such as CO,
layer-by-Iayer growth mode can be extended for many times (Li et al.,
1994b).
To form sharp interface, one needs to limit interfacial diffusion. The ideal
case is that overlayer and substrate metals are immiscible in alloy phase
diagraph, such as Fe and Cu. It should be noted, however, that even for these
immiscible systems, interdiffusion often occurs within the first several
monolayers around the interface region. For Fe on Cu (100), interdiffusion is
limited to one or two mono layers at room temperature but becomes significant
at a higher temperature. It is important to monitor interdiffusion and reduce it
with lower growth temperature and/or higher growth rate.
Magnetism in Ultrathin Films and Beyond 61

Fe Thickness (ML)
o 4 8 12 16

II III
~

:::i
~
.~
C
<U
.5

"Dirty "Fe/Cu (100)


Fe/Cu(100)
Ts=310K o 5 10 15 20 25 30 35
o 1000 2000 Thickness (ML)
Deposition time (s)
(a) (b)

Figure 2.3 RHEED oscillations of Fe growth on Cu (100). (a) clean CLi et al. ,
1994a) and Cb) "dirty" with surfactant CLi et al. , 1994b).

2.2.2 Wires

Besides fabricating 2D films, growth techniques are increasingly explored to


self-assemble lateral structures such as wires and dots. Collective, long-range
interactions can cause meso- to nano-scale self-organization with lateral
compositional, morphological and structural modulations to further lower the
symmetry of the system. Self-assembly not only offers an intrinsically simpler,
faster and less costly alternative route to lithography, but also promises
thermodynamically stable structures even beyond the lithographic limits. The
fabrication of quasi-1 D nanostriped systems has been a big challenge until
recently, when real progress was made. Adatoms diffuse on a substrate
surface and stick to step edges to achieve step-flow growth. Steps provide a
template for forming stripes/wires in this growth mode. Before a full
monolayer is formed, one obtains epitaxial monolayer (MU stripes. Analogous
to the three fi 1m-growth modes, there is row-by-row growth, island formation
and one or more wetting rows followed by island formation (Himpsel et al. ,
1998). Such row-by-row growth introduces additional oscillations in reflected
beam intensities as lateral periodicity also contributes to interference of the
coherent beams. Specifically, an additional half-monolayer peak appears for
row-by-row growth on a stepped surface with a terrace width distribution,
which can be used to characterize such a growth mode (Gambardella et al. ,
2000). Step decoration has been observed for Fe on W (110) (Eimers et al. ,
62 Dongqi Li

1994), Fe/Cu (111) (Shen et al., 1997), Fe/Pd(110) (Li et al., 2001;
Roldan et al. , 2001) , etc. with stripe widths from one atomic row to '" 102 A.
It is known that the initial growth of metals on Pd (110) surfaces is highly
anisotropic and tends to form nanoscale stripes even on flat surfaces (Bucher
et al. , 1994). By using a stepped Pd (110) substrate with the steps along the
fast diffusing direction, i. e., < lTo >, smooth, straight Fe stripes are
encouraged.

2.2.3 Dots

The same mechanism that causes film roughening can potentially be used as a
self-assembly route to create ordered arrays of dots or wires in uniform sizes.
For a system with lattice mismatch, both the surface energies Y, and interface
energy Yi are strain dependent and may vary with film thickness. A very
common way to release strain is via misfit dislocations, especially when defect
density is high. With such a strain relieved mechanism, Y, and Yi can be
virtually thickness independent, and the film remains smooth. For large misfit,
however, a surface roughening is favored, which allows a partial relaxation of
the strain by purely elastic deformation of the film and substrate. In this case,
an SK growth mode with one or more wetting layers and subsequent 3D growth
should occur. Such self-assembled quantum dots have been realized and
extensively investigated in semiconductor systems (Politi et al. , 2000), such
as Ge/Si, GeSi/Si and InAs/GaAs, where stress-driven SK growth mode
results in dots of narrow size distributions and sometimes in relatively ordered
arrays (Floro et aI., 1998). Most recently, it has also been observed in a
metallic system: Co on Ru(OOOl). For epitaxial growth of Co on Ru(OOOl) up
to several hundreds of nanometers, either 3D islands (dots) or holes in flat film
networks (antidots in smooth films) may form. As seen in Fig. 2. 4a and b,
these dots and antidots have well-defined shape, surprisingly smooth surfaces,
and relatively narrow size distribution (Yu et al., 2001 a). From surface
energy considerations, Co should wet Ru to form a smooth film, given the Co
surface energy of 2.709 J/m 2 and Ru of 3.409 J/m 2 (Bauer and van der Merwe,
1986). In addition, there is a large lattice mismatch of '" 8% for Co/Ru, which
favors a SK growth mode. Interestingly, dots and antidots are equivalent in
this model in relieving strain, with antidots being favored for thicker films
(Tersoff and Le Goues, 1994). While good lattice match has been one of the
major criteria to guide epitaxial growth, it becomes possible that well-chosen
lattice mismatched systems could be utilized to fabricate strain-engineered
regular magnetic nanostructure arrays with different sizes and periodicity.
A general challenge in self-assembly is associated with placement and
Magnetism in Ultrathin Films and Beyond 63

nm
12
6
o
100
200
300
nm

(a) (b)

40
20
o 200 400 600 800 nm
o
(c) (d)

(e)

Figure 2.4 Typical AFM images of Co Ca) dot, Cb) antidot, Cc) dot chain, Cd) stripes,
Ce) dot array CYu et al. , 2001a, 2001b; Li and Yu, 2002a).

alignment of the structures at predetermined locations, a task that is routine for


lithographic fabrication. For example, magnetic patterned media consisting of
a magnetic dot array would require dot alignment along tracks. It would be
even more demanding to produce complex spintronic devices (Hathaway and
Prinz, 1995) by means of self-assembly. As seen in Fig. 2. 4c- e, we have
aligned self-assembled magnetic dots, stripes from connecting dots and dot
arrays along substrate grooves that form from residual polishing scratches/step
bunching. This offers the promise that magnetic dot arrays or arbitrary
arrangements could be fabricated by self-assembly with the assistance of
lithographic substrate patterning.
64 Dongqi Li

2. 3 Magnetic Properties in Low Dimensional Systems

2.3. 1 Metastable Structures


When the size of a system becomes small enough, metastable structural
phases can become energetically possible. For example, fcc phase is a high-
temperature phase for bulk Fe, which can be stabilized at room temperature
through epitaxial growth on lattice-matched (0 = 3. 59 A for fcc Fe) fcc
substrates such as Cu (0 = 3. 61 A) (Thomassen et al., 1992; Li et al. ,
1994a, 1994b) and diamond (0 =3.57 A) (Li et al. , 1998a, 1998b). On the
other hand, fcc Fe is also magnetically metastable, with multiple magnetic
states such as ferromagnetic (FM) and antiferromagnetic (AF), depending on
atomic volume and distortion. The interplay of these instabilities makes it the
richest magnetic film system thus far. Figure 2. 2 shows an experimental
magnetic phase diagram, which indicates that depending on film thickness and
deposition temperature, the film exhibits slightly different structural phases and
dramatically different magnetic phases (Li et al. , 1994a). The phase of 6 -
11 ML, marked as II (a) in Fig. 2.5, is particularly interesting since it has a
ferromagnetic surface and antiferromagnetic bulk as illustrated in Fig. 2. 6a.
The total magnetization, as measured with polar Kerr signal, oscillates as a
function of film thickness. Unlike the normal systems where surfaces tend to be
"soft" with phase transition temperature lower than that of the bulk, the Curie
temperature ( T c) of the fcc Fe surface is at"" 250 K, higher than the Neel
temperature ( TN) of the bulk at"" 200 K or less. Such a surface ferromagnetic
order is supported by a structural distortion at the surface, which results in a
face-center-tetrahedral (FCT) surface and a more isotropic fcc bulk. Indeed,
the fct Fe phases in regions I and II (b) (Fig. 2. 5) are ferromagnetic. It
should be noted that such a TN is still significantly larger than that of the fcc Fe
precipitants in a Cu matrix, indicating a subtly different phase. This AF phase
has been suggested (Li et al., 1994a) and recently confirmed (Qian et al. ,
2001) as a spin-density wave state with a period of "" 2. 6 ML. The
experimental temperature dependence of the magnetization is understood using
a self-consistent local mean-field theory as shown in Fig. 2. 6b (Camley and Li,
2000). This is the first example of a thin antiferromagnet with a low TN being
stabilized by a neighboring ferromagnet with a higher T c.
In summary, magnetism is intimately connected to structure since inter-
atomic exchange integrals of the electrons and the central ingredient of
magnetic exchange interactions are highly sensitive to the distance and
symmetry of the neighboring electrons. Nanostructures make it possible to
Magnetism in Ultrathin Films and Beyond 65

350
II (a)

300 : [F(s)lAF(b)]~
(fcc)
.-

g
h~
250
~

III

II (b)
200
FI F//
(fet) (bee)
150
0 4 8 12 16
Fe thickness (ML)

Figure 2.5 Experimental magnetic phase diagram of fcc Fe/Cu(001) (Li et al. , 1994).

(a)

6
~

;:; 5
~
~

"<1)E 4
0
E 3
M(T) --11 layers
... ..
+..
u
.~
" 2
+
bJl
oj
E
"E
~
0
50 100 300

(b)

Figure 2.6 (a) Schematics of the antiferromagnetic phase; (b) experimental (dots)
and theoretical (lines) temperature dependence of magnetization (Li et al., 1994).

create metastable or frustrated structures and therefore new magnetic


materials.
66 Dongqi Li

2.3.2 Dimensionality and Phase Transition


As a physical system shrinks in one or more directions, its thermodynamic
properties can become qual itatively different as the system changes its
dimensionality. The transitions among different dimensions occur in finite size
systems which are also intrinsically interesting. For example, in a finite size
system, the divergence of correlation length can be limited by the physical
size of the system. This results in, among other things, a lower phase
transition temperature for smaller size systems following a finite size scaling
relationship:
(2. 1)

where d is the characteristic length and A is the shift exponent that describes
the scaling behavior (Fisher and Ferdinard, 1967). Such a phenomenon was
observed in magnetic thin films, where T e decreases with film thickness
following a 30-to-20 scaling. Most recently, 20-to-10 scaling has also been
seen in step-decorated magnetic nanowires such as Fe wires on W ( 110)
(Eimers et aI., 1994) and Fe wires on Pd(110) (Li et al., 2001).
Submonolayer Fe wedges were grown on a stepped Pd ( 110) substrate.
RHEEO suggests that Fe atoms decorate the steps to form nanostripes, as
illustrated in Fig. 2. 7a. These stripes are ferromagnetic above 0.3 ML Fe
coverage or -6 A average stripe width, and have a magnetic easy axis along
the surface normal. As shown in Fig. 2. 7b, the T c ' defined as the onset of
magnetization, initially increases rapidly as a function of coverage, or stripe
width, and saturates at -0.8-1 ML. As shown in Fig. 2. 7b, the Teexhibits
finite-size scaling with a shift exponent A of 1.2 0.3, consistent with two-
dimensional (20) Ising expectations, which yields A = 1 as 20 stripes change
their width (Fisher and Ferdinand, 1967; Fisher and Barber, 1972).
In addition to the T e shift effect, it is also anticipated that the phase
transition of a finite system should be rounded, such that

5=IlT cc1 / d o, (2.2)


Teo
where the broadened transition region 11 T is also due to the fact that the
correlation length cannot diverge in a finite system width (Fisher and
Ferdinand, 1967). For a finite-size 20 Ising system, e = 1, indicating a
significant broadening in the transition region as d decreases. While this
scaling behavior is equally as fundamental as the T e shift, it is far less
discussed and has not been positively observed experimentally. It would be
interesting in the future to further understand the novel phenomena in
nanomagnetism in this content. For example, the magnetization M (T) of the
Fe wires on Pd( 110) exhibits, instead of the normal sharp drop at T e' unusual
Magnetism in Ultrathin Films and Beyond 67

Figure 2.7 Ca) Schematics of step decoration; Cb) finite size scaling of the Curie
temperature for Fe stripes CLi et al. , 2001),

exponential decay with increasing temperature over a large temperature of 40


- 120 K (li et al. , 2001), It may be related to the broadened transition for a
low-dimensional system,

2.3.3 Surface/Interface Electronic Structure


The two fundamental elements of magnetism-magnetic moment and exchange
coupl ing-are both determined by the electronic structure of the system, In a
magnetic solid with itinerant electrons such as Fe, Ni and Co, their magnetic
moment is always smaller than that of the corresponding isolated atoms, where
the electrons are completely local ized. In the meantime, the interatomic
overlap of the itinerant electron wave functions in a solid gives rise to
exchange coupling among atoms. Surfaces and interfaces, due to the reduced
coordination numbers compared to that of bulk, is in some sense between the
isolated atoms and bulk of a solid. The electrons are often more localized at
surfaces/interfaces compared to that of the bulk and may form specific
surface/ interface state or resonance. (The difference of the two is that the
former locates in a bulk energy gap while the later does not). Such an altered
electronic structure has strong impact on magnetic properties of surface/
interface and is critical for thin films and nanostructures. For example, it is
68 Dongqi Li

known that magnetic moments at surfaces, as a consequence of electron


local ization, are often enhanced compared to these of the bulk (Freeman and
Wu, 1991>. Magnetocrystalline anisotropy, which is determined by the weak
spin-orbit coupl ing, is strongly altered at surfaces/ interfaces to give rise to
surface/ interface magnetic anisotropy (Wang et al. , 1993; Wu and Freeman,
1999). At interfaces, the electronic states of overlayer and substrate also
hybridize and form new states with the characteristics of both. This is the
fundamental basis of magnetic proximity, where some nearly magnetic
elements, such as Pd and Pt, can obtain a small, induced moment when they
are adjacent to magnetic metals.
One example of magnetic surface state is seen in Gd (Li et al., 1991),
where a highly localized, majority spin, 5d surface state appears near the
Fermi level (Fig. 2.8). It results in an enhanced surface moment (Wu et al. ,
1991), even though the 5d electrons only contribute to a fraction of the total
moment. The moment mainly comes from the fully localized 4f electrons,
which provide 7 J1s The spin character of the surface state indicates that the
surface and bulk are ferromagnetically coupled. An example of the correlation
between electronic structure and magnetic properties is shown clearly with
hydrogen adsorption of Gd surface (Li et ai., 1993), where the surface

Gd(OOOI)
Normal emission
hv =32.7 eV

90K

300K

4 321
Binding energy(eV)

Figure 2. 8 Spin polarized photoemission spectra at different temperatures at


normal emission. The majority-and minority-spin bulk bands are marked with up and
down arrows. The majority-spin peak near the Fermi level is the magnetic surface
state (Li et al. , 1995a).
Magnetism in Ultrathin Films and Beyond 69

electrons become more itinerant and non-magnetic. The most dramatic


magnetic property of Gd is the enhanced Curie temperature at the surface
(Weller et al. , 1985; Tang et aI., 1993), which is still under debate (Arnold
and Pappas, 2000). The relative degree of electron itinerancy also affects
finite temperature magnetism (Nolting et aI., 1994). While the 5d bulk band
behaves mainly as Stoner model, where exchange splitting decreases with
increasing temperature (Li et aI., 1992, 1995a; Bode et aI., 1999; Donath
et aI., 1996), the more localized surface state shows more of spin-mixing
behavior (Li et al. 1992, 1995a; Bode et al. , 1998), where spin-polarization
of the spin-polarized states decreases at higher temperature.

2. 3. 4 Quantum Size Effects


When a nanostructure is well ordered and clean, the electrons can remain
phase coherent at the length scale of its physical size. The walls of the
nanostructure would therefore confine the electrons to form quantum well states
just as the classical textbook example of particle-in-a-box in quantum
mechanics. A good example is the giant magnetoresistance (GMR)
multi layers , which have generated interests in great applications for magnetic
information and sensing technologies as well as vigorous fundamental
research. In these ferromagnetic/nonferromagnetic multi layers , the energetics
of the magnetic quantum-well (OW) states in the nonferromagnetic spacer
component mediate an interlayer magnetic coupling that oscillates in sign as a
function of spacer thickness (Ortega et al., 1993; Bruno, 1993; van
Schilfgaarde and Harrison, 1993). These materials exhibit GMR when
antiferromagnetically coupled as an applied magnetic field restores the
ferromagnetic al ignment.
The OW states are determined by two factors: the spatial confinement
and the periodicity of the electronic states. First, in metallic multi layers , the
energy offsets of bands of the same symmetry of the two adjacent metals can
confine the electronic states. If the exchange splitting of the magnetic metal
provides the energy offset, then the spin-dependent potential barrier results in
confinement of only one spin sub-band and, hence, spin-polarized OW states,
as seen in Fig. 2. 8. The existence of these latter states affects both the
coupl ing strength (Mathon et al., 1993) and GMR ( Mathon, 1991). Ortega
et al. ( 1993) were first to use electron spectroscopy to probe coupled-layer
systems and observed sp-band derived OW states at the Brillouin zone (BZ)
center in Cu overlayers on Co( 100). Figure 2. 9 demonstrates that at the thick
end of a Co wedge the 2-MLCu OW states are fully formed and, thus, the
confinement is maximal, while at the thin end the intensity has decayed
because the states have leaked out across the Co barrier and merged with
those of the bulk Cu substrate. The d- and sp-OW states of the Cu over layer
70 Dongqi Li

leak across the Co barrier at different rates. The characteristic decay lengths
are (0. 60. 2)A for the d-OW state and (2. 20. 6)A for the sp-state. The
results can be understood in terms of electron tunneling through a finite
potential barrier (Li et al. , 1995b). Note that the characteristic length of the
sp-OW state coincides with the interfacial Co thickness needed to optimize the
GMR of permalloy ICu multi layers (Parkin, 1993; Li et al. 1995b). This
highlights the importance of establishing an effective barrier to promote spin-
dependent scattering in order to optimize the GMR.

Figure 2.9 Spin-polarized photoemission spectra of 2 ML of Cu on thick fcc Co(001)


grown epitaxial onto a Cu ( 100) single crystal. The solid (open) symbols are for
majority (minority) spin. Note that the OW states are of minority spin character (Li et
al. , 1995b).

Second, the periodicity with which these magnetic OW states cross the
Fermi level E F correlates as a function of thickness with the periodicity of the
oscillatory interlayer magnetic coupling (Ortega et ai., 1993; Stiles, 1993;
Garrison et ai., 1993; Smith et al., 1994). The coupling periodicity is
determined solely by an extremal spanning vector of the spacer's Fermi
surface. Equivalently, the physics can be understood based on a static spin
density wave forming in the spacer due to the Ruderman-Kittel-Kasuya-Yosida
(RKKY) interaction (Stiles, 1993; van Schilfgaarde and Harrison, 1993;
Koell ing, 1994). Both descriptions (RKKY and OW) correlate electronic
structure with magnetic properties and predict the same periodicity, though
the former is at perturbation limit and is more problematic for quantitative
calculations of coupling strength. Figure 2. 10 shows that the intensities of the
OW states sweeping through E F at k II = O. 72 O. 05,X oscillate with a period
Magnetism in Ultrathin Films and Beyond 71

of (17 2) A, which coincides with the 18-A long period of the oscillatory
magnetic coupling of the Fe/Cr system. Later theoretical investigations
suggest that this OW state has great intensity, though not spin-polarized.
Nevertheless, the results demonstrate that angle-resolved photoemission
provides a novel methodology to search k-space for the features responsible
for the interlayer magnetic coupling.

o
.o-~--c:r ~ O-<J""':~o_::::o~_=~::-. _
..
,.J:T
po. 0

I/ 0

0
1
I
I
I
I
I
I
I
I
/
I

Figure 2. 10 Peak intensities of the sp-QW state at 1.6 eV (solid circles) and d-QW
state at 2.4 eV (open circles, normalized to the same saturation intensity as the sp-
state). Inset: Schematic of the wedge-shaped sample (Li et ai., 1995b).

2.3.5 Domains and Domain Walls

The characteristic length scales of magnetic domains and their interactions,


-- 10' - 10 3 nm, are relatively large compared with those of other sized-
related magnetic phenomena and are therefore among the first phenomena
investigated in small magnetic structures. As a structure is small enough,
i. e. , at sub-micron range, it becomes energetically unfavorable to include
domain walls in multi-domain states. This is why magnetic thin films are often
single-domain. For magnetic dots, even non-uniform configurations become
possible. Depending on thickness, lateral size and magnetic anisotropy, the
ground state of a dot can be either a single-domain state or a vortex state with
magnetization circling in the dot and tilting away from the surface at the
center. Single-domain is favored for smaller dots as exchange interaction
becomes dominant. The self-assembled Co dots discussed previously show
single domain with in-plane magnetization in their virgin states. Micromagnetic
calculations indicate that they are metastable at> 120 nm. Around the phase
boundary between vortex and single-domain states, the two states are almost
72 Dongqi Li

bi-stable, which can irreversibly be converted to each other under an external


field (Yu et al. , 2002).
For the dot chains grown on grooved substrates, the magnetization of the
dots within one chain prefers to line up along the chain direction (Fig.2. 11),
which indicates the existence of significant dipolar-like interactions among the
dots in the chain, as well as a uniaxial magnetic anisotropy with easy axis
along the chains. Such a uniaxial anisotropy is confirmed with azimuthal
dependence of remanent magnetization.

Figure 2.11 The intensities at E F vs. Cr thickness at kll =0. nrx,


which yield a
periodicity of (17 2) A. Independent thickness sequences are indicated with
different symbols (Li et aI., 1997).

Inter-dot magnetic correlation along a dot chain can be quantified with the
probabilities of nearest-neighbor dots with parallel and anti-parallel moments,
P tt and P I I ' and the deducted pair-correlation function, < SiS i+ 1 > =

: tt ~: tl , as plotted in Fig. 2. 12 as a function of their center-to-center


tt II
distance (Li et al. 2002b). Parallel pairs of nearest neighbors greatly out-
number the anti-parallel ones, confirming the ferromagnetic coupling among
dots. Such a system can be modeled with a classical 10 Ising chain at zero
fields. The pair correlation function of the spins at sites i and i + 1 in such a
model system is exactly solved as (Ising, 1925; Thompson, 1972)

EII-E tt )
<Si S i+l> = tanh ({3J) = tanh ( 2k T
s
' (2.3)

where J is the coupling constant and {3 is 1/(k s n. The solid line in the inset
of Fig. 2. 12 indicates the fitting result with dipolar inter-dot interaction. This fit
indicates that the 10 Ising model describes the basic physics of the magnetic
Co dot chains.
It is well known that an infinite 10 Ising chain does not have long-range
order. Our data confirm, however, that ferromagnetic correlation indeed
Magnetism in Ultrathin Films and Beyond 73

(a)

(b)
Figure 2. 12 ea) AFM and eb) the corresponding MFM images of one magnetic dot-
chain. Note that the moments of the dots tend to point to the same direction, indicating a
ferromagnetic coupling among the dots in a chain eli et al. , 2002b).

exists at short range even in a 10 system presumably because ferromagnetic


coupling is present. Within a finite length, the chains appear to be
magnetically ordered with sizable total magnetization, as seen in Fig. 2. 11.
Since all quasi-1D systems have finite length, this in itself acts to break the
symmetry and permits the ordering.
1.0

J 0.8

25% ""'~- 0.6


"o
.~ 0.4
]
20% 8 0.2
...
~ 0
.2 15% -0.2 !:-:::--~:-;o--~-;;----;;-;!
200 400 600 800
:0
Pair distance (nm)
'2"
.n
Q..
10%

5%

0
200 700 800

Figure 2. 13 Probability of the nearest neighbor dots epairs) in parallel eA) or anti-
parallel e.) alignment as a function of pair disctance. In the inset, the diamond symbols
e.) shows the experimental pair correlation function derived from the probability.
74 Dongqi Li

2.4 Conclusions

Nanomagnetism is a rich and rapidly developing area. Its richness resides in


the existence of multiple characteristic lengths and their interactions, which
lead to exciting a new physics. Its development strongly depends on the
progress in state-of-art fabrication techniques for well-defined, highly ordered
individual structures, arrays, and hybrid structures. The goal is to design and
produce tailored functional nanomaterials for new applications.

References
Arnold C. S. and D. P. Pappas. Phys. Rev. Lett. 85: 5202 (2000)
Bauer, E. and Z. Kristallogr. 110: 372(1958)
Bauer, E. and J. H. van der Merwe. Phys. Rev. B 33: 3657(1986)
Bode, M. et al. Appl. Phys. A 66: S 121 (1998)
Bode, M. et al. Phys. Rev. Lett. 83: 3017 (1999)
Bruno, P. J. Magn. Magn. Mater. 121: 248 (1993)
Bucher, J. -P., E. Hahn, P. Fernandez, C. Massobrio and K. Kern.
Europhysics Lett. 27: 473 (1994)
Camley, R. E. and Dongqi Li. Phys. Rev. Lett. 84: 4709 (2000)
Donath, M. , B. Gubanka and F. Passek. Phys. Rev. Lett. 77: 5138 (1996)
Egelhoff, W. F. Jr. and I. Jacob. Phys. Rev. Lett. 62: 921(1989)
Eimers, H. J. , J. Hauschild, H. Hache, U. Gradmann, H. Bethge, D. Heuer
and U. Kohler. Phys. Rev. Lett. 73: 898 (1994)
Fisher, M. E. and A. E. Ferdinand. Phys. Rev. Lett. 19: 169 (1967)
Fisher, M.E. and M.N. Barber. Phys. Rev. Lett. 28: 1516(1972)
Floro, J. A. , E. Chason, M. B. Sinclair, L. B. Freund and G. A. Lucadamo.
Appl. Phys. Lett. 73: 951 (1998)
Freeman, A. J. and R. Q. Wu. J. Magn. Magn. Mater. 100: 497 (1991)
Fullerton, E. E., C. H. Sowers, J. P. Pearson, X.Z. Wu, D. Lederman and
S.D. Bader. Appl. Phys. Lett. 69: 2438 (1996)
Gambardella, P. , M. Blanc, H. Brune, K. Kuhnke and K. Kern. 61: 2254
(2000)
Garrison,K., Y. ChangandP. D. Johnson. Phys. Rev. Lett. 71: 2801
(1993)
Hathaway, K. and G. Prinz. Phys. Today 48: 24 (1995)
Himpsel, F. J. , J. E. Ortega, G. J. Mankey and R. F. Willis. Advances in
Physics 47: 511 (1998)
Ising, E. Z. Phys. 31: 253 (1925)
Koelling, D. D. Phys. Rev. B 50: 273 (1994)
Magnetism in Ultrathin Films and Beyond 75

Li, D.O., C. W. Hutchings, P. A. Dowben, C. Hwang, R. -T. Wu, M.


Oneil ion , A.B. Andrews and J.L. Erskine. J. Magn. Magn. Mater. 99: 85
( 1991)
Li, D. 0., J.D. Zhang, P. A. Dowben and M. Onellion. Phys. Rev. B 45:
7272 (1992)
Li, D. 0., J. D. Zhang, P. A. Dowben and M. Onellion. Phys. Rev. B 48:
5612 (1993)
Li, D. 0., M. Freitag, J. Pearson, Z. O. Oiu and S. D. Bader. Phys. Rev.
Lett. 72: 3112 (1994a)
Li, D. 0., M. Freitag, J. Pearson, Z. O. Oiu and S. D. Bader. J. Appl.
Phys. 76: 6425 (1994b)
Li, D. O. , P. A. Dowben, J. E. Ortega and F. J. Himpsel. Phys. Rev. B
49: 7734 (1994c)
Li, D. O. , J. Pearson, S. D. Bader, D. N. Mcilroy, C. Waldfried and P.
A. Dowben. Phys. Rev. B (Rapid Commun.) 51: 13,895 (1995a)
Li, D. 0., J. Pearson, P. D. Johnson, J. E. Mattson and S. D. Bader.
Phys. Rev. B 51: 7195 (1995b)
Li, D. 0., J. Pearson, S. D. Bader, E. Vescovo, D.-J. Huang, P. D.
Johnson and B. Heinrich. Phys. Rev. Lett. 78: 1154 (1997)
Li, D. 0., D. J. Keavney, J. Pearson, S. D. Bader, J. Pege and W.
Keune, Phys. Rev. B 57: 10044 (1998a)
Li, D. 0., D. J. Keavney, J. Pearson, J. S. Jiang, S. D. Bader and W.
Keune. J. Vac. Sci. Tech. A 16: 2326 (1998b)
Li, D.O., B. Roldan C., J. Pearson, S. D. Bader and W. Keune. Phys.
Rev. B 64: 144,410 (2001)
Li, D.O. and C. T. Yu. MRS Proceedings. in press (2002a)
Li, D.O., C.T. Yu, J. Pearson and S.D. Bader. submitted to PRL (2002b)
Mathon, J. J. Magn. Magn. Mater. 100: 527 (1991)
Mathon, J. , M. Villeret, D. M. Edwards and R. B. Muniz. J. Magn. Magn.
Mater. 121: 242 (1993)
Nolting, W. , T. Dambeck and G. Borstel. Z. Phys. B 94: 409 (1994)
Ortega, J. E. , F. J. Himpsel, G. J. Mankey and R. F. Willis. Phys. Rev. B
47: 1540 (1993)
Parkin, S. S. P. Phys. Rev. Lett. 71: 1641 (1993)
Politi, P. , G. Grenet, A. Amarty, A. Ponchet, J. Villain. Phys. Reports
324: 271 (2000)
Prinz, G. J. Mag. Mag. Mater. 200: 57 (1999)
Oian, D. , X. F. Jin, J. Barthel, M. Klaua and J. Kirschner, Phys. Rev.
Lett. 87: 227,204 (2001)
Roldan, B. , J. Pearson, C. T. Yu, D. O. Li and S. D. Bader. J. Vac. Sci.
Technol. A 19: 1182 (2001)
Sander, D. , A. Enders and J. Kirschner. J. Mag. Mag. Mater. 200: 439
(1999)
Shen, J. , R. Skomski, M. Klaua, H. Jenniches, S. Sundar Manoharan and
76 Dongqi Li

J. Kirschner. Phys. Rev. B 56: 2340 (1997)


Smith, N. V., N. B. Brookes, Y. Chang and P. D. Johnson. Phys. Rev.
B 49: 332 (1994)
Stiles, M. D. Phys. Rev. B 48: 7238 (1993)
Tang, H., D. Weller, T.G. Walker, J.C. Scott, C. Chappert, H. Hopster,
A. W. Pang, D. S. Dessau and D. P. Pappas. Phys. Rev. Lett. 71: 444
( 1993)
Tersoff, J. and F. K. Le Goues. Phys. Rev. Lett. 72: 3570 (1994)
Thomassen, J. , F. May, B. Feldmann, M. Wuttig and H. Ibach. Phys. Rev.
Lett. 69: 3831C 1992)
Thompson, C. J. Phase Transitions and Critical Phenomena 1: 177 ( 1972)
van Schilfgaarde, M. and W. A. Harrision. Phys. Rev. Lett. 71: 3870
(1993)
Wang, D. S. , R. O. Wu, A. J. Freeman. Phys. Rev. Lett. 70: 869 (1993)
Weller, D., S. F. Alvarado, W. Gudat, K. Schroder and M. Campagna.
Phys. Rev. Lett. 54: 1555 (1985)
Wu, R. 0., C. Li, A. J. Freeman and C. L. Fu. Phys. Rev. B44: 9400
( 1991)
Wu, R.O. and A.J. Freeman. J. Magn. Magn. Mater. 200: 498 (1999)
Wuttig, M. , B. Feldmann, J. Thomassen, F. May, H. Zillgen, A. Brodde,
H. Hannemann and H. Neddermeyer. Surf. Sci. 291: 14 (1993)
Yu, C. T. , Dongqi Li, J. Pearson and S. D. Bader. Appl. Phys. Lett. 78:
1228 (2001a)
Yu, C. T. , Dongqi Li, J. Pearson and S. D. Bader. Appl. Phys. Lett. 79:
3848 (2001b)
Yu, C. T. , J. Pearson and D. O. Li. J. Appl. Phys. in press (2002)

The work supported by DOE BES-MS under # W-31- 109-ENG-38.


3 Classical and Quantum Magnetization Reversal
Studied in Nanometer-Sized Particles and Clusters

Wolfgang Wernsdorfer

3. 1 Introduction

Nanometer-sized magnetic particles have generated continuous interest since


the late 1940s because the study of their properties has proved to be
scientifically and technologically very challenging. In particular, it was
recognized that the ferromagnetic state, with a given orientation of the particle
moment, has a remanent magnetization if the particle is small enough. This
was the starting point of huge permanent magnets and magnetic recording
industries. However, despite intense activity during the last few decades, the
difficulties in making nanoparticles of good enough quality have slowed the
advancement of this field. As a consequence, for 50 years, these appl ications
concentrated above and then near the micrometer scale. In the last decade,
this has no longer been the case because of the emergence of new fabrication
techniques that have led to the possibility of making small objects with the
required structural and chemical qualities. In order to study these objects, new
techniques such as magnetic force microscopy, magnetometry based on micro-
Hall probes or micro-SQUIDs (superconducting quantum interference device)
were developed. This led to a new understanding of the magnetic behavior of
nanoparticles, which is now very important for the development of new
fundamental theories of magnetism and in modeling new magnetic materials for
permanent magnets or high density recording.
In order to put this chapter into perspective, let us consider Fig. 3. 1,
which presents a scale of size ranging from macroscopic down to nanoscopic
sizes. The unit of this scale is the number of magnetic moments in a magnetic
system (roughly corresponding to the number of atoms). At macroscopic
sizes, a magnetic system is described by magnetic domains (Hubert and
Schafer, 1998) that are separated by domain walls. Magnetization reversal
occurs via nucleation, propagation, and annihilation of domain walls (see the
hysteresis loop on the left in Fig. 3. 1 which was measured on an individual
Cozr particle of 1 IJm x O. 8 IJm and a thickness of 50 nm). Shape and
elliptic CoZr
width of domain walls depend on the material of the magnetic system, on its
78 Wolfgang Wernsdorfer

size, shape and surface, and on its temperature (Aharoni, 1996). The
material dependence of the domain walls has motivated the definition of two
length scales: CD the {) defined by {j
domain wall width {j {) J A I/ K
= v' ~the
and (2)the
= ,JA/
exchange length i\A defined by i\A = ./AIMsM s where A is the exchange energy,
K is the crystalline anisotropy constant, and M s is the spontaneous
magnetization. Qualitatively, the first definition shows that anisotropy energy
favors a thin wall, while the exchange energy favors a thick wall. For very
small crystalline anisotropy, the first definition suggests an infinite domain wall
width which has a large total energy. This is due to the magnetostatic energy
term that can be reduced by subdividing the ferromagnetic crystal into
domains. Therefore, for very small crystall ine anisotropy, the domain wall
width is of the order of magnitude of the exchange length i\.A. Both length scales
can range from submicrometer scales in alloys to atomic scales in rare earth
systems.
Mesoscopic physics
Macroscopic _..
1--------
_00-- -----<._
1.. _ Nanoscopic

Permanent Micron Nanopm1icles


Nanoparticles Clusters Molecular Individual
magnets pal1icles
particles Clusters Spins

I I I I I I I I I
S =10 20
8=10 10 10 10 8 106 105 104 103 102 10 1

Multi-domain Single-domain Magnetic Moments

Nucleation, propagation and Uniform rotation Resonant tunneling, quantization,


annihilation of domain walls curling quantum thermodynamics

-II--J------
-II-~--

-40 -20 0 20 40
-1
-I t:::::'
-100
::::::::::::==--------'
t:::::====~--=-=-- 0
c7::--
100 -I o
~oH(mT)
poH(mT) ~oH(mT)
poH(mT) ~oH(mT)
poH(mT)
(a) (b) (c)

Figure 3. 1 Scale of size that goes from macroscopic down to nanoscopic sizes. The hysteresis
loops are typical examples of magnetization reversal via nucleation, propagation and annihilation
of domain walls (a), via uniform rotation (b), and via quantum tunneling (c).

When the system size is of the order of magnitude of {j{) or i\, A, the
formation of domain walls requires too much energy. Therefore, the
magnetization remains in the so-called single-domain state. Hence, the
magnetization might reverse by uniform rotation, curling or other nonuniform
modes (see hysteresis loop in the middle of Fig. 3. 1). In this chapter we
discuss mainly this size range where the physics is rather simple (Sections 3. 3
and 3. 4).
For system sizes well below {j{) and i\,A, one must take into account
Classical and Quantum Magnetization Reversal ....
.. 79

explicitly the magnetic moments (spins) and their couplings. The theoretical
description is complicated by the particle's boundaries (Freemann and Wu,
1991; Pastor et al.,
al. , 1995; Kohl and Bertsch, 1999).
At the smallest size (below which one must consider individual atoms and
spins) there are either free clusters made of several atoms (Apsel et al. ,
1996; Billas et ai.,
aI., 1997) or molecular clusters which are macromolecules
with a central complex containing magnetic atoms. In the last case,
measurements on the Mn12 acetate and FesFee molecular clusters showed that the
physics can be described by a collective moment of spin S = 10 =
(Section 3. 5. 1). By means of simple hysteresis loop measurements, the
quantum character of these molecules showed up in well-defined steps which
are due to resonance quantum tunneling between energy levels (see hysteresis
loop on the right in Fig. 3. 1).
In the following sections, we review the most important theories and
experimental results concerning the magnetization reversal of single-domain
particles and clusters. Special emphasis is put on single-particle
measurements avoiding complications due to distributions of particle size,
shape, etc. Measurements on particle assemblies have been reviewed in
al., (1997). We mainly discuss the low-temperature regime in
Dormann et ai.,
order to avoid spin excitations.
In Section 3. 2, we briefly review the commonly used measuring
techniques. Among them, electrical transport measurements, Hall probes,
and micro-SQUID techniques seem to be the most convenient techniques for
low-temperature measurements. Section 3.3 discusses the mechanisms of
magnetization reversal in single-domain particles at zero Kelvin. The influence
of temperature on the magnetization reversal is reported in Section 3. 4.
Finally, Section 3. 5 shows that for very small systems or very low
temperature, magnetization can be reversed via tunneling.

3.2 Single-Particle Measurement Techniques

The following sections review commonly used single-particle measuring


techniques avoiding complications due to distributions of particle size, shape
etc., which are always present in particle assemblies (Dormann et al. ,
1997). Special emphasis is laid on the micro-SQUID technique which allowed
us the most detailed studies at low temperatures.

3.2.1 Overview of Single-Particle Measurement Techniques


The dream of measuring the magnetization reversal of an individual magnetic
80 Wolfgang Wernsdorfer

particle goes back to the pioneering work of Neel (1949a, 1949b). The first
realization was published by Morrish and Yu (1956). These authors employed
a quartz-fiber torsion balance to perform magnetic measurements on individual
micrometer-sized y-Fey -Fe22 0
033 particles. With their technique, they wanted to
avoid the complication of particle assemblies which are due to different
orientations of the particle's easy axis of magnetization and particle-particle
dipolar interaction. They aimed to show the existence of a single-domain state
in a magnetic particle. Later on, other groups tried to study single particles,
but the experimental precision did not allow a detailed study. A first
breakthrough came via the work of Knowles (1978) who developed a simple
optical method for measuring the switching field, defined as the minimum
applied
appl ied field required to reverse the magnetization of a particle. However, the
work of Knowles failed to provide quantitative information on well-defined
particles. More recently, insights into the magnetic properties of individual
and isolated particles were obtained with the help of electron holography
onomura et a!.,
(Tonomura
(T aI., 1986), vibrating reed magnetometry (R ichter , 1989),
(Richter,
Lorentz microscopy (Hefferman et a!. a!. , 1991), magneto-
al. , 1991; Salling et al.
optical Kerr effect (Bardou et a!., al., 1996) and magnetic force microscopy
(Chang and Chu, 1994; Ledermann et al., 1994). Recently, magnetic
nanostructures have been studied by the technique of magnetic linear dichroism
in the angular distribution of photoelectrons or by photoemission electron
microscopy (Bansmann et al., a!., 1999; Meiweis-Broer, 1999). In addition to
magnetic domain observations, element-specific information is available via
the characteristic absorption levels or threshold photoemission. We refer to
the literature concerning other domain observation techniques (Hubert and
Schafer, 1998). Among all mentioned techniques, most of the studies have
been carried out using magnetic force microscopy at room temperature. This
technique has an excellent spatial resolution, but time-dependent
measurements are difficult due to the sample-tip interaction.
Only a few groups were able to study the magnetization reversal of
individual nanoparticles or nanowires at low temperatures. The first
magnetization measurements of individual single-domain nanoparticles and
nanowires at very low temperatures were presented by Wernsdorfer et al.
( 1995a). The detector (an Nb microbridge-DC-SQUID) and the studied
particles were fabricated using electron-beam lithography. Coppinger et al.
( 1995) investigated the magnetic properties of nanoparticles by resistance
measurements. They observed the two-level fluctuations in the conductance
which they attributed to fluctuations of magnetization of self-organizing ErAs
quantum wires and dots in a semi-insulating GaAs matrix. By measuring the
electrical resistance of isolated Ni wires with diameters between 20 and
40 nm, Giordano and Hong studied the motion of magnetic domain walls (Hong
and Giordano, 1995; Wegrowe et al., a!., 1998). Other low-temperature
...
Classical and Quantum Magnetization Reversal ... 81

techniques that may be adapted to single-particle measurements are Hall


probe magnetometry (Kent et al., 1994; Lok et al., 1998; Schweinbock et
al. , 2000), magnetometry based on magnetoresistance (Gros et al., 1997;
Gallagher et aI.,
ai., 1997; Wegrowe et ai.,
aI., 1999) or spin-dependent tunneling
with Coulomb blockade (Schelp et aI.,
al., 1997; Gueron et ai.,
aI., 1999). At the
time of this writing, the micro-SQUID technique allows the most detailed study
of the magnetization reversal of nanometer-sized particles (Wernsdorfer et
al. , 1996a, 1997b, 1997c; Bonet et aI.,
al., 1999; Jamet et al. , 2001). The
following section reviews the basic ideas of the micro-SQUID technique.

3.2.2 Micro-SQUID Magnetometry

SQUIDs have been used very successfully for magnetometry and voltage or
current measurements in the fields of medicine, metrology and science. They
are mostly fabricated from a Nb-AIO x -Nb trilayer, several hundreds of
nanometers thick. The two Josephson junctions are planar tunnel junctions with
an area of at least O. 5 I-lm IJm22 . In order to avoid flux pinning in the
superconducting film, the SQUID is placed in a magnetically shielded
environment. The sample's flux is transferred via a superconducting pick up
coil to the input coil of the SQUID. Such a device is widely used, as the signal
can be measured by simple lock-in techniques. However, this kind of SQUID is
not well suited to measuring the magnetization of single submicron-sized
samples as the separation of SQUID and pickup coil leads to a relatively small
coupling factor. A much better coupling factor can be achieved by coupling the
sample directly with the SQUID loop. In this arrangement, the main difficulty
arises from the fact that the magnetic field applied to the sample is also
applied to the SQUID. The lack of sensitivity to a high field applied in the
SQUID plane, and the desired low temperature range led to the development
of the micro-bridge-DC-SQUID technique (Wernsdorfer, 1996) which allows us
to apply several teslas in the plane of the SQUID without dramatically reducing
the SQUID's sensitivity.
The planar Nb DC can be constructed by using standard electron beam
lithography, and the magnetic particle is directly placed on the SQUID loop
(Fig. 3.2) (Wernsdorfer et aI.,ai., 1995a, 1995c). The SQUID detects the flux
through its loop produced by the sample magnetization. For hysteresis loop
measurements, the external field is appl applied
ied in the plane of the SQUID, so that
the SQUID is only sensitive to the flux induced by the stray field of the sample
magnetization. Due to the close proximity between sample and SQUID,
magnetization reversals corresponding to 10 3 IJB g. , the
/.is can be detected, e. g.,
magnetic moment of a Co nanoparticle with a diameter of 2 - 3 nm.
82 Wolfgang Wernsdorfer

Particle

" ~//
Josephson junctions

3. 2 Drawing of a planar Nb DC on which a ferromagnetic particle is placed.


Figure 3.2
The SQUID detects the flux through its loop produced by the sample magnetization.
Due to the close proximity between sample and SQUID a very efficient and direct flux
coupling is achieved.

3. 3 Mechanisms of Magnetization Reversal at Zero Kelvin

As already briefly discussed in the introduction, for a sufficiently small


magnetic sample it is energetically unfavorable to form a stable magnetic
domain wall. The specimen then behaves as a single magnetic domain. For
the smallest single-domain particles, the magnetization is expected to reverse
by uniform rotation of magnetization (Section 3. 3. 1). For somewhat larger
ones, nonuniform reversal modes are more likely, for example, the curling
reversal mode (Section 3. 3. 2. 1 ). For larger particles, magnetization
reversal occurs via a domain wall nucleation process starting in a rather small
volume of the particle. For even larger particles, the nucleated domain wall
can be stable for certain fields. The magnetization reversal happens then via
nucleation and annihilation processes (Section 3. 3. 2.3).
2. 3). In these sections we
neglect temperature and quantum effects.
The following section discusses in detail the uniform rotation mode that is
used in many theories, in particular in Neel, Brown and Coffey's theory of
magnetization reversal by thermal activation (Section 3. 4) and in the theory of
macroscopic quantum tunneling of magnetization (Section 3. 5).

3.3. 1 Magnetization Reversal by Uniform Rotation (Stoner-


Wohlfarth Model)
The model of uniform rotation of magnetization, (Stoner-Wohlfarth model)
developed by Stoner and Wohlfarth (1948) and Neel (1947),
(1947>, is the simplest
classical model describing magnetization reversal. One considers a particle of
an ideal magnetic material where exchange energy holds all spins tightly
Classical and Quantum Magnetization Reversal . . . 83

parallel to each other, and the magnetization magnitude does not depend on
the space. In this case the exchange energy is constant, and it plays no role in
the energy minimization. Consequently, there is competition only between the
anisotropy energy of the particle and the effect of the applied field. The
original study by Stoner and Wohlfarth assumed only uniaxial shape
anisotropy, which is the anisotropy of the magnetostatic energy of the sample
induced by its nonspherical shape. Thiaville has generalized this model for an
arbitrary effective anisotropy which includes any magnetocrystalline anisotropy
and even surface anisotropy (Thiaville, 1998, 2000).
In the simplest case of uniaxial anisotropy, the energy of a Stoner-
Wohlfarth particle is given by
E = KVsin 2 I/> - iJo (I/> - 8)
/-10 M VHcos (
s (3.1)
where KV is the uniaxial anisotropy energy which depends on the shape of the
particle, V is the volume of the particle, M ss is the magnetization of
saturation, H is the magnitude of the appl ied field, and
applied I/> and 8 are the angles
of magnetization and applied
appl ied field respectively, with respect to the easy axis
of magnetization. The potential energy of Eq. (3. 1) has two minima separated
by an energy barrier. For given values of 8 and H, the magnetization lies at
an angle I/> which locally minimizes the energy. This position can by found by
I/> of Eq. (3.1):
equating to zero the first derivative with respect to dE/dl/> =0. The
(3. 1): dE/d
second derivative provides the criteria for maxima and minima. The
magnetization reversal is defined by the minimal field value at which the
energy barrier between the metastable minimum and the stable one vanishes;
that is, at dE/dl/>
dE/d = = d2E/d2
E/d1/>2 == O. A short analysis yields the angular
dependence of this field, called the switching field H~w' H~w, or in dimensionless
units:

(3.2)

where H a == 2K / (iJo
(/-10 M s ) is the anisotropy field. The angular dependence of
h~w is plotted in Fig. 3. 3.
Contrary to h~w' the hysteresis loops cannot be expressed analytically
and have to be calculated numerically. The result is seen in Fig. 3. 4 showing
the component of magnetization projected along the direction of the applied
field; that is, M H = M s cos (I/> ( - 8). Such loops are often called Stoner-
Wohlfarth hysteresis loops. It is important to note that single-particle
measurement techniques do not measure this component M M H.
H For example,

for the micro-SQUID technique, with the easy axis of magnetization in the
plane of the SQUID and perpendicular to the current direction in the SQUID
wire (Fig. 3.2), one measures a magnetic flux that is proportional to M s sin (I/
()
(Fig. 3.5).
84 Wolfgang Wernsdorfer

90'
1.0 0 0
120'
120 60'
60

l------~O'

240' 300'
0
270'
270

Figure 3.3 Angular dependence of the Stoner-Wohlfarth switching field h~w = H~w/ H~w /
H a [Eq. (3. 2)]. This curve is often called the
the""Stoner-Wohlfarth
Stoner-Wohlfarth astroid. " Cases (1)
to (3) correspond to Eqs. (3. 5) - (3. 9) concerning the field dependence of the
anisotropy barrier height.

0'
0
0

[0 0

_[
-I ~~~::::..::..::..::..=......=:.;..:.;.;.;.:.:...--.J
Fs~<!t=:::=::::...:::..::..=......=:.:.:.:.:.:.:.:...--.J

-1.5 -1.0 -0.5 0 0.5 1.0 1.5


HI H a

Figure 3.4 Hysteresis loops of a Stoner-Wohlfarth particle for different field angles
9. The component of magnetization along the direction of the applied field is plotted;
(J.
that is, MH=M
MH=M,s cos (cp-(J).
(1/1-9).

The main advantage of this classical theory is that it is sufficiently simple


to add some extra features to it, as presented in the following.

3.3.1.11
3.3.1. Generalization of the Stoner-Wohlfarth Model
The original model of Stoner and Wohlfarth assumed only uniaxial shape
anisotropy with one anisotropy constant; that is, one second-order term, see
Eq. (3. 1). This is sufficient to describe highly symmetric cases like a prolate
spheroid of revolution or an infinite cylinder. However, real systems are often
quite complex, and the anisotropy is a sum of mainly shape (magnetostatic) ,
magnetocrystalline,
magnetocrystall ine, magnetoelastic and surface anisotropy. One additional
Classical and Quantum Magnetization Reversal ....
.. 85

....-----.....
..................... ...
~~ - _ _ 1 " - _ _ j
------."'.-.~-___r-_1

F-.-J--.....",'"'-'
F---L--~.~--.~--=----._ ..... ....

Figure 3.5 Hysteresis loops of a nanocrystalline elliptic Co particle of 70 nm x 50 nm x


25 nm. The dashed line is the prediction of the Stoner-Wohlfarth model of uniform rotation
of magnetization. The deviations are due to nonuniform magnetization states.

complication arises because the different contributions of anisotropies are often


aligned in an arbitrary way one with respect to each other. All these facts
motivated a generalization of the Stoner-Wohlfarth model for an arbitrary
effective anisotropy which was done by Thiaville (1998, 2000).
Similar to the Stoner-Wohlfarth model, one supposes that the exchange
interaction in the cluster couples all the spins strongly together to form a giant
spin whose direction is described by the unit vector m. The only degrees of
freedom of the particle's magnetization are the two angles of orientation of m.
The reversal of the magnetization is described by the potential energy
E(m, H) =
= Eo(m) - lJo
IJo VMsmH (3.3)

where V and M ss are the magnetic volume and the saturation magnetization of
the particle respectively, H is the external magnetic field, and Eo (m) is the
magnetic anisotropy energy which is given by
(3.4)

Eshape(m) is the magnetostatic energy related to the cluster shape. EMC(m) is


the magnetocrystalline anisotropy (Me) arising from the coupling of the
magnetization with the crystalline lattice, similar as in bulk. Esurface(m) is due
to the symmetry breaking and surface strains. In addition, if the particle
experiences an external stress, the volumic relaxation inside the particle
induces a magnetoelastic (ME) anisotropy energy E ME ME (m). All these
developed in a power series of m~ m~ m; with p =
bedeveloped
anisotropy energies can be
a+ b + c = = 2, 4, 6,
6,,,,
... giving the order of the anisotropy term. Shape
86 Wolfgang Wernsdorfer

anisotropy can be written as a biaxial anisotropy with two second-order terms.


Magnetocrystalline anisotropy is in most cases either uniaxial (hexagonal
systems) or cubic, yielding mainly second- and fourth-order terms. Finally, in
the simplest case, surface and magnetoelastic anisotropies are of second
order.
Thiavi lie proposed a geometrical method to calculate the particle's
Thiaville
energy and to determine the switching field for all angles of the applied
magnetic field yielding the critical surface of switching fields which is
analogous to the Stoner-Wohlfarth astroid (Fig. 3. 3).
The main interest of Thiaville' s calculation is that measuring the critical
surface of the switching field allows one to find the effective anisotropy of the
nanoparticle. The Knowledge
knowledge of the latter is important for temperature-
dependent studies (Section 3.4) and quantum tunneling investigations
(Section 3.5).
3. 5). Knowing precisely the particle's shape and the
crystallographic axis allows one to determine the different contributions to the
effective anisotropy.
Thiaville's calculation predicts also the field dependence of the energy
/lE close to switching [E = ( 1- H / H~w) 1] which is important
barrier height flE
to know for temperature-dependent studies (Sections 3. 4 and 3. 5). Three
cases as follow have to be distinguished.
( 1) In the majority of cases except the two following cases (see case (1)
(1)
in Fig. 3.3),

AE
LJ.::=:::::: V2
4KV - -cos
-E Y 3/2 -- E oE 3/2 (3.5)
3 $
where KV is the anisotropy energy constant, y is the angle of incidence
between the local normal to the critical surface and the field sweeping
direction, and p is the radius of curvature of the focal curve (Thiaville, 2000)
at H~w. It is important to emphasize that all these variables can be found
experimentally by measuring the critical surface of the switching field. For
uniaxial anisotropy-that is, the 20 2D Stoner-Wohlfarth case-Eq. (3.5) (3. 5 )
becomes
2 3/2 I tel't e 1 1/3
/3
/lE ~ 4KV ( - )
flE::=::::::4KV(-) co 3/2 (3.6)
3 1 + I cote 112/32
/3 E

e
where e is the angle of the applied field with respect to the easy axis of
magnetization (Eq. (3. 1)).
(2) At glancing incidence (see case (2) in Fig. 3.3) with respect to the
(y=n/2), the power of E is different, yielding
critical surface (Y=TT/2),
/lE~E~E3 (3.7)

E~~ has been calculated only in the two-dimensional case (Thiaville,


where E
1998).
= TT/2
(3) At a cusp point where y = n/2 (see case (3) in Fig. 3.3),
3. 3),
Classical and Quantum Magnetization Reversal ... 87

(3.8)
where E; has been calculated only in the two-dimensional case (Thiaville,
1998). In the case of uniaxial anisotropy, that is the 2D Stoner-Wohlfarth
case, Eq. <3.
(3. 8) becomes
(3.9)
<3.9)

where H.=H. = 2 2~K (Eq.(3.1. e = 0 andlT/2,


(Eq. <3. 1. This equation is only valid for e=o and 11'/2,
Po M s
110
and it is valid for O~H~H . This famous result of Neel (1949a, 1949b) has
often been wrongly used for assemblies
assembl ies of nanoparticles where it is very
e
difficult to achieve the conditions = = 0 and 11'/2
IT/2 (Victora, 1989). until now,
only the power 3/2 and 2 (cases (1) and (3 has been found by single-
particle measurements (Wernsdorfer et aI., 1997b, 1997c) (see also
Figs. 4.3 and 4.4).
Another simple analytical approximation for the field dependence of the
energy barrier t:.E(H) was derived numerically by Pfeiffer (1990a, 1990b):
= KV(
t:.E = H/H~w)O
KV(l1 - H/ = KVE o
H~w)O = (3.10)
0.10)
a = O. 86 + 1. 14
where 0=0.86+ 14h~w'
h~w, and h~w is given by Eq. (3. 2). This approximation
is good for the intermediate field regime, that is, for fields H not too close to
H. and not too small.

3.3.1.2 Experimental Evidence for Magnetization Reversal by Uniform Rotation


In order to demonstrate experimentally the uniform rotation mode, the angular
dependence of the magnetization reversal has often been studied (Aharoni,
1996). However, a comparison of theory with experiment is difficult because
magnetic particles often have a nonuniform magnetization state that is due to
rather compl icated shapes and surfaces, crystall ine defects, and surface
anisotropy. In general, for many particle shapes the demagnetization fields
inside the particles are nonuniform, leading to nonuniform magnetization states
(Aharoni, 1996). An example is presented in Fig. Fig.3.
3. 5 which compares typical
hysteresis loop measurements of an ell iptical Co particle, fabricated by
elliptical
electron beam lithography, with the prediction of the Stoner-Wohlfarth model.
Before magnetization reversal, the magnetization decreases more strongly
than predicted because the magnetic configuration is not collinear
coli inear as in the
Stoner-Wohlfarth model, but instead presents deviations mainly near the
particle surface. The angular dependence of the switching field agrees with the
Stoner-Wohlfarth model only for angles e? e ~ 30 where nonlinearities and
defects playa less important role (Wernsdorfer et aI.,al. , 1995a, 1995b).
Studies of magnetization reversal processes in ultrathin magnetic dots
with in-plane uniaxial anisotropy showed also switching fields that are very
close to the Stoner-Wohlfarth model, although magnetic relaxation experiments
clearly showed that nucleation volumes are by far smaller than an individual dot
volume (Fruchart et al., 1999). These studies show clearly that switching
88 Wolfgang Wernsdorfer

field measurements as a function of the angles of the applied field cannot be


taken unambiguously as proof of a Stoner-Wohlfarth reversal.
The first clear demonstration of the uniform reversal mode has been found
with Co nanoparticles (Wernsdorfer et al., 1997b) and BaFeO nanoparticles
(Wernsdorfer et aI., 1997c), the latter having a dominant uniaxial
magnetocrystalline anisotropy. The three-dimensional angular dependence of
the switching field measured on BaFeO particles of about 20 nm could be
explained with the Stoner-Wohlfarth model taking into account the shape
crystal Iine anisotropy of BaFeO (Bonet et al.,
anisotropy and hexagonal crystalline al. ,
1999). This explanation is supported by temperature- and time-dependent
measurements yielding activation volumes which are very close to the particle
volume (Section 3.4).
We present here the first measurements on individual cobalt clusters of
3 nm in diameter containing about a thousand atoms (Figs. 3.6 and 3.7)
(Jamet et al. , 2001). In order to achieve the needed sensitivity, Co clusters
preformed in the gas phase are directly embedded in a co-deposited thin Nb
film that is subsequently used to pattern micro-SQUIDs. A laser vaporization
and inert gas condensation source is used to produce an intense supersonic
beam of nanosized Co clusters which can be deposited in various matrices
under ultra-high-vacuum (UHV) conditions. Due to the low-energy deposition
regime, clusters do not fragment upon impact on the substrate (Perez et al. ,
1997). The niobium matrix is simultaneously deposited from a UHV electron
gun evaporator leading to continuous films with a low concentration of
embedded Co clusters (Jamet et al. , 2000). These films are used to pattern
planar microbridge-DC-SQUIDs by electron beam lithography. The later ones
allow us to detect the magnetization reversal of a single Co cluster for an
applied magnetic field in any direction and in the temperature range between
O. 03 and 30 K (Section 3. 2. 2). However, the desired sensitivity is only
achieved for Co clusters embedded into the microbridges where the magnetic
flux coupling is high enough. Due to the low concentration of embedded Co
clusters, we have a maximum of 5 non interacting particles in a microbridge
which is 300 nm long and 50 nm wide. We can separately detect the
magnetization switching for each cluster. Indeed they are clearly different in
intensity and orientation because of the random distribution of the easy
magnetization directions. The cold mode method in combination with the blind
method (Wernsdorfer, 1996) allows us to detect separately the magnetic
signal for each cluster.
High-resolution transmission electron microscopy observations showed
that the Co clusters are well-crystallized in a face-centered cubic (fcc).
structure (Fig. 3.6) with a sharp size distribution (Jamet et al. , 2000). They
mainly form truncated octahedrons (Fig. 3. 7) (Jamet et al. , 2001).
3.7)
Figure 3. 8 displays a typical measurement of switching fields in three
dimensions of a 3 nm Co cluster at T = = 35 mK. This surface is a three-
dimensional picture directly related to the anisotropy involved in the
Classical and Quantum Magnetization Reversal ...
... 89

Figure 3. 6 High-resolution transmission electron microscopy observation along a


[110J direction of a 3 nm cobalt cluster exhibiting an fcc structure.

Figure 3.7 Scheme of a typical cluster shape with light-gray atoms belonging to
the 1289 atoms truncated octahedron basis and dark-gray atoms belonging to the
(111) and (00 1) added facets.
(001)

magnetization reversal of the particle (Section


CSection 3. 3. 1). It can be reasonably
CSection 3. 3. 1. 1). We
fitted with the generalized Stoner - Wohlfarth model (Section
obtain the following anisotropy energy
EoCm)/V =- KIm;
Eo(m)/V K 1 m; K2m~
+K 2 m; - K4(m~.m~.
K 4 Cm; m~ m~m;
+ m; m; + m~m;).
m~ m;).
(3.11)

where K 1 and K 2 are the anisotropy constants along z and x, x. the easy and
hard magnetization axes,
axes. respectively. K 4 is the fourth-order anisotropy
Cx' y' z') coordinate system is deduced from (xyz)
constant and the (x' Cxyz) by a 45
=2. 2 xX 105 J/m3 , K 22 =0.
rotation around the z axis. We obtained K 11 =2.2 =O. 9 X 105 J/m3
and K 4 == O. 1 X 105 J/m3 . The corresponding theoretical surface is shown in
90 Wolfgang Wernsdorfer

Fig. 3. 9. Furthermore, we measured the temperature dependence of the


switching field distribution (Section 3. 4. 3. 2). We deduced the blocking
~ 14 K and the number of magnetic atoms in this
temperature of the particle T B ""'"
~ 1500 atoms (Section 3. 4. 3. 2). Detailed measurements on
particle: N ""'"
about 20 different particles showed similar three-dimensional switching field
distributions with comparable anisotropy.

JioH:
f.1o H z(T)
(T)
0.3

0.4

Figure 3.8 Side view of the experimental three-dimensional angular dependence of


the switching field of a 3 nm Co cluster at 35 mK. This surface is symmetrical with
respect to the H,
H x -- Hy
H y plane, and only the upper part (Hz>
(Hz>O)
0) is shown. Continuous
lines on the surface are contour lines on which Hz is constant.

JioH: (T)
0.3

0.4

Figure 3.9 Side view of the theoretical switching field surface considering second- and
fourth-order terms in the anisotropy energy.

In the following, we analyze various contributions to the anisotropy energy


of the Co clusters. Fine structural studies using extended X-ray absorption fine
structure (EXAFS) measurements (Jamet et al. , 2000) were performed on
500 nm thick niobium films containing a very low concentration of cobalt
clusters. They showed that niobium atoms penetrate the cluster surface to
almost two atomic monolayers because cobalt and niobium are miscible
elements. Further magnetic measurements (Jamet et al. , 2000) on the same
samples showed that these two atomic monolayers are magnetically dead. For
this reason, we estimated the shape anisotropy of the typical nearly spherical
deposited cluster in Fig. 3. 7 after removing two atomic monolayers from the
surface. By calculating all the dipolar interactions inside the particle, and
assuming a bulk magnetic moment of J.1at /-I at = 1. 7 J.1B'
/-IB' we estimated the shape
K 1 ~O. 3 x
constants: K1"""'0.
anisotropy'constants:
anisotropy 3
X 105 J/m along the easy magnetization axis
Classical and Quantum Magnetization Reversal ... 91

and K2~0. 1 x 10 5 J/m3 along the hard magnetization axis. These values are
much smaller than the measured ones which means that E Eshape
shape is not the main
cause of the second-order anisotropy in the cluster.
The fourth-order term K 4 = O. 1 X 10 5 J/m3 should arise from the cubic
magnetocrystalline anisotropy in the fcc cobalt clusters. However, this value is
smaller than the values reported in previous works (Lee et al. , 1990; Chuang
et al. , 1994). This might by due to the different atomic environment of the
surface atoms with respect to that of bulk fcc - Co. Taking the value of the
bulk = 1. 2 X 10 J/m only for
5 3
bulk (Lee et al. , 1990; Chuang et aI., al. , 1994), K bulk
the core atoms in the cluster, we find K Me ~ O. 2 X 105 J/m3 , which is in
reasonable agreement with our measurements.
We expect that the contribution of the magnetoelastic anisotropy energy
K ME coming from the matrix-induced stress on the particle is also small.
Indeed, using the co-deposition technique, niobium atoms cover the cobalt
uniformly,, creating an isotropic distribution of stresses. In addition,
cluster uniformly
they can relax preferably inside the matrix and not in the particle volume
because niobium is less rigid than cobalt. We believe therefore that only
interface anisotropy K surtace
surface can account for the experimentally observed
second-order anisotropy terms. Niobium atoms at the cluster surface might
enhance this interface anisotropy through surface strains and magnetoelastic
coupling. This emphasizes the dominant role of the surface in nanosized
systems.
In conclusion, the three-dimensional switching field measurements of
individual clusters give access to their magnetic anisotropy energy. A
quantitative understanding of the latter is still difficult, but it seems that the
cluster-matrix interface provides the main contribution to the magnetic
anisotropy. Such interfacial effects could be promising to control the magnetic
anisotropy in small particles in order to increase their blocking temperature up
to the required range for applications.
3.3. 1. 3 Uniform Rotation Mode with Cubic Anisotropy
We have seen in the previous section that the magnetic anisotropy is often
dominated by second-order anisotropy terms. However, for nearly symmetric
shapes, fourth-order terms (for example, the fourth-order terms of fcc
magnetocrystalline anisotropy) can be comparable or even dominant with
respect to the second-order terms. Therefore, it is interesting to discuss
further the features of fourth order terms. We restrict the discussion to the 20
problem (Chang, 1991; Thiaville, 1998) (see Ref. (Thiaville, 2000) for
3D) .
30)
The reversal of the magnetization is described by Eq. (3.O. 3) that can be
rewritten in 20
E(8) = E o (8) J..lo VMs(Hxcos8
- IJo + H ysin8)
y (3.12)
where V and M s are the magnetic volume and the saturation magnetization of
92 Wolfgang Wernsdorfer

the particle, respectively, e is the angle between the magnetization direction


and x, H x and H yare the components of the external magnetic field along x
and y, and Eo ( e) is the magnetic anisotropy energy. The conditions of
critical fields (dE/de = 0 and d2 E / de 2
2
= 0) yield a parametric form of the
locus of switching fields
2
1(. dE o
0 Eo)
d E O)
(3. 13)
=- 2J.1
H =-
xx VM Sine
2110o VMs
s sine
de ++cose
cose de2

22
1 ( dE o . d Eo)Eo ) (3.14)
(3. 14)
= ++ 2J.1o
Hy = 2110 VMs cose de - sine de 22 ..

As an example we study a system with uniaxial shape anisotropy and cubic


anisotropy. The total magnetic anisotropy energy can be described by
Eo(e) = VK 1 sin 2 (e + eo) + VK 2 sin 2 ecos 2 e (3.15)

where K, and K 2 are anisotropy constants (K, (K 1 could be a shape anisotropy


and K 2 the cubic crystalline anisotropy of an fcc crystal), eo is a constant
which allows one anisotropy contribution to turn with respect to the other one.
Figure 3. 10 displays an example of a critical curve which can easily be
calculated from Eqs. (3. 13) - (3. 15). When comparing the standard Stoner-
Wohlfarth astroid in Fig. 3.3
3. 3 with Fig. 3. 10, we realize that the critical curve
can cross itself several times. In this case, the switching field of magnetization
depends on the path followed by the applied field. In order to understand this
point, let us follow the energy potential (Eqs. (3. 12) and (3. 15)) when
sweeping the applied field as indicated in Fig. 3. 11. When the field is in A,
the energy E has two minima and the magnetization is in the metastable
potential well. As the field increases, the metastable well becomes less and
A-8,-C-O-E,
less stable. Let us compare two paths, one going along A-B,-C-D-E,
the other over 8 B 2 instead of 8,.
B,. Figure 3. 12 shows E in the vicinity of the

Figure 3.10 Angular dependence of the switching field obtained from Eqs. (3.13)-

(3. 15) with K, >0 and K 2 = - K 1. The field is normalized by the factor 2K , . h x =
2 K,.
3 lJo M s5
1J0
H x /(2K,/lJ oM s ),
)' and h yy =H y/(2K'/lJ ).
y /(2K,/lJ o M ss ).
Classical and Quantum Magnetization Reversal ..
... 93

metastable well for different field values along the considered pathsCthe stable
potential well is not presented). One can realize that the state of the
magnetization in C depends on the path followed by the field: Going over B 1
leads to the magnetization state in the left metastable well (1), whereas going
over B 2 leads to the right metastable well (2). The latter path leads to
magnetization switching in 0, and the former one leads to a switching in E.
Note that a small magnetization switch happens when reaching B 1 or B 22 . Point
JI is a supercritical bifurcation.

1.0

Figure 3.11 Enlargement of angular dependence of the switching field of Fig. 3. 3.10.
10.
Two possible paths of the applied field are indicated: Starting from point A and going
over the point B8 11 leads to magnetization reversal in E, whereas going over the point
B
8 2 leads to reversal in D.

2 1I

c~

3. 12 Scheme of the potential energy near the metastable state for different applied
Figure 3.12
fields as indicated in Fig. 3. 11. The balls represent the state of the magnetization, and the
arrows locate the appearing or disappearing well.
94 Wolfgang Wernsdorfer

3. 3. 1. 4 Experimental Evidence for Uniform Rotation Mode with Cubic


Anisotropy
The study of uniform rotation for particles with dominating cubic anisotropy
revealed to be very difficult because nearly all structural defects in particles
lead to dominating uniaxial anisotropy. Nevertheless, we could find few
particles which were sufficiently "perfect" in order to show clearly a field path
dependence of the switching field (Fig. 3. 10) which is the important signature
of strong higher order terms in the potential energy (Eq. (3. 3.3) ). The first
measurement of such a field path dependence of switching fields were
performed on single-domain FeCu nanoparticles of about 15 nm with a cubic
crystalline anisotropy and a small arbitrarily oriented shape anisotropy (Bonet
et aI., 1998). Here we report on the magnetic study of individual single-
domain cobalt nanoparticles which are synthesized by an arc-discharge method
(Guerret-Piecourt et al., 1994). Most of the nanocrystals have a spherical
shape with a diameter ranging from 5 to 30 nm (Fig. 3. 13). All the particles
are encapsulated in a carbon envelope which provides a very efficient
protection against oxidation.

Figure 3. 13 TEM image of an assembly of typical cobalt nanoparlicles encapsulated in


a few graphitic sheets and a large amount of amorphous carbon. Most of the particles
have a spherical shape checked through tilting experiments and frequently contain planar
defects (see arrows) .

The crystallites are pure cobalt and have an fcc structure according to the
electron diffraction patterns obtained from assemblies of particles. The
nanoparticles are mostly single-crystals and often contain twin boundaries and
stacking faults (Fig. 3. 13) since these planar defects are known to occur
frequently in fcc-Co.
In order to place one nanoparticle on the SQUID detector, we disperse
Classical and Quantum Magnetization Reversal ... 95

the particles in ethanol by ultrasonication. Then we place a drop of this liquid


on a chip of about one hundred SQUIDs. When the drop is dry the
nanoparticles stick on the chip due to Van der Waals forces. Only in the case
when a nanoparticle falls on a micro-bridge of the SQUID loop, the flux
coupling between SQUI!? loop and nanoparticle is strong enough for our
measurements. Finally we determine the exact position and shape of the
nanoparticles by scanning electron microscopy.
The measured angular dependence of switching fields of nearly all Co
nanoparticles revealed a dominating uniaxial magnetic anisotropy similar to
previous measurements (Wernsdorfer et al. , 1997b). We estimated the shape
anisotropy of the typical nearly spherical nanoparticles (Fig. 3. 13) and found
that the shape anisotropy constants should be of the order of magnitude of
J/ m3 , i. e. one order of magnitude smaller than the value of the bulk fcc
10 4 J/m
cobalt (Lee et al. , 1990; Chuang et al. , 1994). This result suggests that twin
boundaries and stacking faults (Fig. 3. 13) strongly alter the cubic crystal
symmetry leading to dominating uniaxial anisotropy.

-0.02 0.13/loH:(T)

lloH,(T)

Figure 3. 14 Top view and side view of the experimental three-dimensional angular
dependence of the switching field of a 15 nm Co particle at 35 mK. This surface is
inversion symmetrical with respect to the origin (H= 0). Continuous lines on the surface
are contour lines on which Hz is constant.
96 Wolfgang Wernsdorfer

Nevertheless, we could find few particles which were sufficiently


"perfect" in order to show a more complex switching field surface and a field
path dependence of the switching field (Fig. 3. 10) which is the important
signature of strong higher order terms in the potential energy (Eq. (3. O. 3 .
One example is presented in Figs. 3.14-3.16.
3. 14 - 3. 16. Note that this surface has two
easy axes. In addition, the hard plan is deformed. Such a surface can, in
principle, be generated by the generalized
general ized Stoner-Wohlfarth model when
taking into account that the different contributions of anisotropies are aligned in
an arbitrary way, one with respect to each other.

0.2

-{).2
-0.2 0.2 iJoHz (T)
/loH,(T)

IS Same data as in Fig. 3. 14. Continuous lines on the surface are contour lines
Figure 3. 15
which are parallel to the SQUID plane (defined by H, = 0) .
Hz =0).

0.1

0.05
i:'
[:;"
.......
'-'
0
~
--{).05
-0.05

-0.1
-{).2
-0.2 --{).1
-0.1 0 0.1 0.2
Hy (T)
lloHy(T)
f.1o

Figure 3. 16 Cut of the angular dependence of the switching field as indicated in


Fig.3. 15 (SQUID plane). A clear field path dependence of the switching field was found
(Figs. 3.10-3.12) which shows a strong influence of the cubic crystalline anisotropy.

In order to understand this point qualitatively, let's come back to the


simple 2D case (Eq. 3. 15) where 8 eo0 is a constant which allows to turn the
uniaxial shape anisotropy with respect to the cubic anisotropy. Figure 3. 17
shows the angular dependence of the switching field for a misalignment of
Classical and Quantum Magnetization Reversal ....
.. 97

9 = 10'
80 = 10 which leads to a strong deformation of the curve in Fig. 3. 10.

Figure 3.17 Angular dependence of the switching field obtained from Eqs. (3. 15)-
800 = 10' .
(3. 17) with the same constants as in Fig. 3. 10 but with 9

3. 3. 2 Nonuniform Magnetization Reversal


We have seen in the previous sections that for extremely small particles,
magnetization should reverse by uniform rotation. For somewhat larger single-
domain particles, nonuniform reversal modes are more likely. The simplest
one is the curling mode that is discussed in the following section.

3.3.2.1
3.3.2. 1 Magnetization Reversal by Curling
The simplest nonuniform reversal mode is the curling reversal mode (Frei
et al. , 1957; Aharoni, 1996). The critical parameter is the exchange length
A==VAl
fi/ M s ' delimiting the region of uniform rotation and curling (A is the
exchange energy). Therefore in the case of the size R> A (R is, for example,
the radius of a cylinder),
cylinder) , magnetization reversal via curling is more favorable.
In the following, we review the analytical result of an ellipsoid of rotation,
which can be applied approximately to most of the shapes of nanoparticles or
nanowires (Aharoni, 1999).
The variation of the switching field with the angle 89 (defined between the
applied field and the long axis of the ellipsoid) is given by Aharoni (1997>
(1997)

(3.16)

x.y = 2N xoy
where a xoy x.y are the demagnetization factors, $' = R/
x.y - k /I $'2, N xoy RI
A, and R is the minor semi-axis of the ellipsoid. The parameter k is a
monotonically decreasing function of the aspect ratio of the ellipsoid. This
function is plotted in Fig. 1 of Aharoni (1986). The smallest and highest value
of k is that for an infinite cyl inder (k =
cylinder = 1.079)
1. 079) and a sphere (k == 1.379),
1. 379) ,
98 Wolfgang Wernsdorfer

respectively.
For a long ellipsoid of rotation, the demagnetization factors are given by
n = = 11 -- 2 NN z
n ~ 1
1
Nz = -2--1
2
n- ~
n (n - 1 + Jn
In(n
In Vn 2
- 1)and
1) and N x = NNy = 2
vnz=1 y

(3.17)
(3. 17)
where n is the ratio of the length to the diameter.
For an infinite cylinder, Eq. (3.16) becomes (Frei et al., 1957; Aharoni,
1996)

HO = Ms h , (l + h,) (3.18)
(3. 18)
sw 2 Jhf + (1 + 2h,)cos2e
where h = -1. 079/S'2.
h,t = 079/8'2. Equation (3. 18) is a good approximation for a very
long ellipsoid of rotation. It is plotted in Fig. 3. 18 for several radi
radiii of an infinite
cylinder.
The case of uniform rotation of magnetization was generalized by Thiaville
to an arbitrary anisotropy energy function and to three dimensions
(Section 3. 3. 1. 1). For the curl ing mode, this generalization
general ization is not possible.
However, approximated calculations were proposed (Ishi i, 1991; Aharoni,
1999; Aharoni, 2000) and micromagnetic simulations were performed (Ferre
et al. , 1997).

.ko'=--+--+--+-+-+l+-+--+--l--+-------==~ 0'
I-E==---I--+--+---+-J-IJ-+--+-+-----jf-----~

-90'

Figure 3. 18 Angular dependence of the switching field of an infinite cylinder for


S. For S' < 1,
several reduced cylinder radii S'. " the switching field is given by the
uniform rotation mode (Section 3.3.
3. 3. 1)
,) .

3.3. 2. 2 Experimental Evidence for Magnetization Reversal by Curling


We report here the first studies of isolated nanoscale wires with diameters
smaller than 100 nm, for which single-domain states could be expected
al.,, 1996a, 1997a). The cyl
(Wernsdorfer et al. indrical geometry, with its large
cylindrical
shape anisotropy, is well-suited for comparison with theory.
...
Classical and Quantum Magnetization Reversal ... 99

Ni wires were produced by filling electrochemically the pores of


commercially available nanoporous track-etched polycarbonate membranes of
thicknesses of 10 IJm. The pore size was chosen in the range of 30 nm to
100 nm (Doudin and Ansermet, 1995; Meier et ai., al. , 1996). In order to place
one wire on the SQUID detector, we dissolved the membrane in chloroform
and put a drop on a chip of some hundreds of SQUIDs. Magnetization
measurements were performed on SQUIDs with a single isolated wire.
The angular dependence of the switching field of wires with 45 nm and
92 nm in diameter are shown in Fig. 3. 19. These measurements are in
quantitative agreement with the curling
curl ing mode (Eq. (3. 16. Nevertheless,
dynamical measurements showed a nucleation volume that is much smaller than
the wire volume (Wernsdorfer et al., a!., 1996a, 1997a) (Section 3.4.3.3).
3. 4. 3. 3) .
Therefore, we believe that the magnetization reversal starts close to curling
instability, but the nucleation happens in a small fraction of the wire only, then
rapidly propagates along the whole sample. This picture is also in agreement
with micromagnetic simulations (Ferre, et ai.,
a!., 1997) and the micromagnetic
model of Braun (1999).
90
1.0

30

::: 0

-30

-90

Figure 3. 19 Angular dependence of the switching field of two Ni wires with a


diameter of 45 nm and 92 nm; that is, S' = 1. 4 and 2. 4. The switching fields are
normalized by 125 mT and 280 mT, respectively.

The angular dependence of the switching field of Ni wires with larger


diameters (270 - 450 nm) were measured at room temperature by Lederman
et al.
a!. (1995). Their results could roughly be explained by the curling mode.
3.3.2.3 Magnetization Reversal by Nucleation and Annihilation of Domain
Walls
For magnetic particles that have at least two dimensions much larger than the
domain wall width, the magnetization reversal may occur via nucleation/
propagation and annihilation of one or several domain walls happening at two
or more applied fields. We focus here on a 30-nm-thick elliptic Co particle
100 Wolfgang Wernsdorfer

defined by electron beam lithography and lift-off techniques out of sputtered


thin films (Wernsdorfer et aI., 1996b). The Co film has a nanocrystalline
structure leading to a magnetically soft material with a coercive field value of
3 mT at 4 K. Therefore we neglected the magnetocrystalline anisotropy. The
nanofabricated particles have an elliptic contour with in-plane dimensions of
300 nm x 200 nm and a thickness of 30 nm.
In order to study the domain structure of our particles, we measure the
angular dependence of hysteresis loops. Figure 3.20
3. 20 shows a typical hysteresis loop
of an individual Co particle. The magnetic field is appl ied in the plane of the
applied
particle. The hysteresis loop is mainly characterized by two magnetization
jumps. Starting from a saturated state, the first jump can be associated with
domain wall nucleation and the second jump can be associated with domain
wall annihilation. During these jumps, the magnetization switches in less than
100 I.JS
j..IS (our time resolution). The reversible central region of the hysteresis
loops is evidence for the motion of the domain wall through the particle.
4

~(
(P)
3
2

~II
~ @,, "" "-
@
!e;~:<X a
'" -I
Li: -1
~
,
~\',
\
\
\',
"
"-
"-

-2
-3
~ __
~-- -- -- (D ,IJ
(]) I
I

-4
-100 -50 a0 50 100
j./{]H(mT)
/loH (mT)
Figure 3. 20 Hysteresis loops at O. 1 K of the elliptic Co particle seen in the inset
which shows the electron micrograph of the wire of a microbridge SQUID with the Co
particle (300 nm x 200 nm x 30 nm). The in-plane field is applied along the long axis of
the particles. The domain wall structure in an elliptical particle is also presented
schematically as proposed by van den Berg. Arrows indicate the spin direction. The
near""" 5 mT might be due to the reversal of the center
two small magnetization jumps near"'"
vortex of the domain structure.

The simplest domain structure, showing such a hysteresis loop, has been
proposed by van den Berg (1987) in zero field and calculated for fields smaller
than the saturation field by Bryant and Suhl (1989). (Note that these models
2D and neglect the domain wall width. Therefore, they can give only a
are 20
qual itative description.) This domain structure has been observed
experimentally on low anisotropy circular thin film disks (100 I.Jm
j..Im in diameter)
(Ruhrig et al. , 1990). In zero field, the
using high-resolution Kerr techniques (ROhrig
particle has a vortex-like
vortex-I ike domain wall as shown in Fig. 3.20.
3. 20. When a magnetic
field is applied, this domain wall is pushed to the border of the particle. For
Classical and Quantum Magnetization Reversal...
Reversal . . . 101

higher fields the domain wall is annihilated and the particle becomes single
domain. The main conditions of the model of the van den Berg model are CD
that the magnetic material is very soft and (2) ~ that the system is two-
dimensional. The first condition is satisfied by our particles because they are
made of a randomly oriented nanocrystallized Co, Co. andare very soft. The
second condition is not quite well satisfied; however,
however. MFM measurements
confirmed this domain structure (Fernandez et al., al.. 1998). Furthermore.
Furthermore, we
obtained similar results for thinner particles (10 nm and 20 nm). More
complicated domain structures as proposed by van den Berg (1987;
Wernsdorfer et aI.,al.. 1996b) may be excluded by the fact that similar Co
particles of length smaller than 200 nm are single domain (Wernsdorfer et al. ,
1995a,
1995a. 1995b).
After studying the magnetization reversal of individual particles,
particles. the
question arises as to how the properties of a macroscopic sample are based on
one-particle properties. In order to answer this question.
question, we fabricated a
sample consisting of 1. 8 x 107 nearly identical elliptic Co particles of about the
same dimensions and material as the individual particle studied above. These
particles are placed on an Si substrate with a spacing of 2 IJm. Because of this
large spacing,
spacing. dipole interactions between particles are negligible. Figure
3. 21 shows the hysteresis loop of the array of Co particles when the field is
applied parallel to the long axis of the particle. This hysteresis loop shows the
same characteristics as the hysteresis loop of one particle (Fig. 3. 20)-that
is,
is. nucleation and annihilation of domain walls. Because of switching field
distributions mainly due to surface defects and a slight distribution of particle
sizes, shapes, and so on.
sizes. shapes. on, the domain wall nucleation and annihilation are no
longer discontinuous,
discontinuous. although they are still irreversible. They take place
along continuous curves with a width of about 10 mT.
1.5

1.0
0(5
~
"0
>.
ME 0.5

.:f-
:f- 0
1:,
b
xX -<l.5
-0.5
~
-1.0
I
-1.5
-120 -80 -40 0 40 80 120
/loR (mT)
J.loH(mT)

3. 21 Hysteresis loops of the magnetic moment of the array of 1. 8 x 10 7 Co


Figure 3.21
particles (300 nm x 200 nm x 30 nm). The in-plane field is applied along the long axis of
the particles.
102 Wolfgang Wernsdorfer

3.
3.44 Influence of Temperature on the Magnetization Reversal

The thermal fluctuations of the magnetic moment of a single-domain


ferromagnetic particle and its decay towards thermal equilibrium were
introduced by Neel (1949a,
(1949a. 1949b) and further developed by Bean and
Livingston (Bean,
(Bean. 1955,
1955. Bean and Livingstone,
Livingstone. 1959) and Brown (1959, (1959.
1963a,
1963a. 1963b). The simplest case is an assembly of independent particles
having no magnetic anisotropy. In the absence of an applied
appl ied magnetic field,
field.
the magnetic moments are randomly oriented. The situation is similar to
paramagnetic atoms where the temperature dependence of the magnetic
susceptibility follows a Curie behavior,
behavior. and the field dependence of
magnetization is described by a Brillouin function. The only difference is that
the magnetic moments of the particles are much larger than those of the
paramagnetic atoms. Therefore.
Therefore, the quantum mechanical Brillouin function can
be replaced by the classical limit for larger magnetic moments.
moments, namely the
Langevin function. This theory is called superparamagnetism. The situation
changes. however. as soon as magnetic anisotropy,
changes, however, anisotropy. which establishes
establ ishes one or
more preferred orientations of the particle's magnetization (Section 3. 3) is
present. In the following,
following. we present an overview of the simplest model
describing thermally activated magnetization reversal of single isolated
nanoparticles which is called the Neel-Brown model. After a brief review of the
model (Section 3. 4. 1 ), ). we present experimental methods to study the
thermally activated magnetization reversal (Section 3. 4. 2). Finally,Finally. we
discuss some applications of the Neel-Brown model (Section 3. 4. 3).

3.4. 1 Neel-Brown Model of Thermally Activated MagIletization


Magnetization Reversal
In the Neel-Brown model of thermally activated magnetization reversal,
reversal. a
magnetic single-domain particle has two equivalent ground states of opposite
magnetization separated by an energy barrier which is due to shape and
crystalline anisotropy. The system can escape from one state to the other by
thermal activation over the barrier. Just as in the Stoner-Wohlfarth model.
model,
they assumed uniform magnetization and uniaxial anisotropy in order to derive
a single relaxation time. Neel supposed further that the energy barrier
between the two equilibrium states is large in comparison to the thermal
energy k B T, T. which justified a discrete orientation approximation (Neel,
(Neel.
1949a,
1949a. 1949b). Brown criticized Neel' s model because the system is not
expl icitly treated as a gyromagnetic one (Brown,
explicitly (Brown. 1959,
1959. 1963a,
1963a. 1963b).
Brown considered the magnetization vector in a particle to wiggle around an
Classical and Quantum Magnetization Reversal ...
... 103

energy minimUm, then jump to the vicinity of the other minimum, then wiggle
around there before jumping again. He supposed that the orientation of the
magnetic moment may be described by a Gilbert equation with a random field
term that is assumed to be white noise. On the basis of these assumptions,
Brown was able to derive a Fokker-Planck equation for the distribution of
magnetization orientations. Brown did not solve his differential equation.
Instead he tried some analytic approximations and an asymptotic expansion for
the case of the field parallel or perpendicular to the easy axis of
magnetization. More recently, Coffey et al. (1995, 1998) found by numerical
methods an exact solution of Brown's
Brown' s differential equation for uniaxial
anisotropy and an arbitrary applied field direction. They also derived an
asymptotic general solution for the case of large energy barriers in comparison
to the thermal energy k BB T. In the following, we consider a simplified version
of the Neel-Brown model and propose three techniques to check the validity of
this model.

3.4.2 Experimental Methods for the Study of the Neel-Brown Model


As discussed in the previous section, in the Neel-Brown model of thermally
activated magnetization reversal, a magnetic single-domain particle has two
equivalent ground states of opposite magnetization separated by an energy
barrier due to, for instance, shape and crystalline anisotropy. The system can
escape from one state to the other either by thermal activation over the barrier
at high temperatures or by quantum tunneling at low temperatures
(Section 3.5).
3. 5). At sufficiently low temperatures and at zero field, the energy
barrier between the two states of opposite magnetization is much too high to
observe an escape process. However, the barrier can be lowered by applying
a magnetic field in the opposite direction to that of the particle' s
magnetization. When the applied field is close enough to the switching field at
zero temperature H~w'
H~w, thermal fluctuations are sufficient to allow the system to
overcome the barrier, and the magnetization is reversed.
In the following, we discuss three different experimental methods for
studying this stochastic escape process. The methods are called waiting time,
switching field and telegraph noise measurements.
3. 4. 2. 1
3.4.2.1 Waiting Time Measurements
The waiting time method corresponds to magnetization relaxation measurements on
assemblies of particles. In this case, the decay of magnetization is often
logarithmic in time, due to the broad distribution of switching fields. For
individual particle studies, waiting time measurements give direct access to
the switching probability (Fig. 3. 22). At a given temperature, the magnetic
w near the switching field H~w. Next,
field H is increased to a waiting field H w
the elapsed time until the magnetization switches is measured. This process is
104 Wolfgang Wernsdorfer

repeated several hundred times, yielding a waiting time histogram. The


integral of this histogram yields the switching probability. Finally, the
switching probability is measured at different waiting fields H w ' and at
different temperatures, in order to explore several barrier heights.
pet)
P(t)
I

H=const.
T=const.

(a)

counts. (Hsw(dH/dt,T)
(Hsw(dH/dt,D)

dH/dt=const.

B'
'8' T=const.

H
(b)

w
M
_T(Hy,T)

Hy=const.
T=const.

o
(c)

Figure 3. 22 Schema of three methods for studying the escape from a potential well:
waiting time and telegraph noise measurements give direct access to the switching time
probability P( t), whereas switching field measurements yield histograms of switching
fields. (a) waiting time measurement, (b) switching field measurement, (c) telegraph
noise measurement.

According to the Neel-Brown model, the switching time P ((t)t) probability


that the magnetization has not switched after a time t is given by
P(
Pet) =
t) = tiT
e- tlT (3.19)
<3.19)
and T (inverse of the switching rate) can be expressed by an Arrhenius law of
the form
= Toe-E(HJ/kBT
T(T,H) = Toe-E(H)/kBT <3.20)
(3.20)

where To
TO is the inverse of the attempt frequency. A simple analytical
approximation for the field dependence of the energy barrier E (H) is
E(H)
E ( H) = Eo
E o (1
(l - H/H~w)a
H / H~w)" = EoE
Eo E"a (3.21)
(3. 21)

where Eo is the energy barrier at zero field, and E is defined as a reduced field
Classical and Quantum Magnetization Reversal.
Reversal . . . 105

difference value. It can be shown that the exponent a is in general equal to


1. 5 (Victora, 1989). Using the Stoner-Wohlfarth analytical expressions of
E ( H) (Stoner and Wohlfarth, 1948), we can show numerically that a is near
E(H)
1. 5 (Wernsdorfer, 1996) and increases up to a value of 2 (more frequently
cited in the literature) if the applied field forms an angle smaller than a few
degrees with the easy axis of magnetization.
3. 4. 2. 2 Switching Field Measurements
For single particle studies, it is often more convenient to study magnetization
reversal by ramping the applied field at a given rate and measuring the field
value as soon as the particle magnetization switches. In this case, the
switching process is characterized by a distribution of switching fields
(Fig. 3.22) evaluated by Kurkijarvi (1972) and Garg (1995). The most
probable switching field H sw is given by

;0 cT
Hsw = H~w { 1 - [ k T In ( VE,,-I )Jl/"}
)JI/"} (3.22)

where c=k Bs H~w/(ToaEo),


H~w/(ToaEo)' and v is the field sweeping rate. The width of
the switching field distribution is given by

~ 0
'""-'
a ----- H sw a
CJ .......
-l(k
-.l(kEs T)I/"[k
T)l/"[kEs TIn (~)J(l-")/"
B B
,,-1
,,-I (3.23)
0 0 VE

The entire switching field distribution P ( H) can be calculated iteratively by


the following equation (Kurkijarvi, 1972):

P(H) = = T-
I
1
v-t f
(H) v-t 1 -
H

o
P(H')dH' J.
J. (3.24)

3.4. 2. 3 Telegraph Noise Measurements


At zero applied field, a single domain magnetic particle has two equivalent
ground states of opposite magnetization separated by an energy barrier. When
the thermal energy k sB T is sufficiently high, the total magnetic moment of the
particle can fluctuate thermally, like a single spin in a paramagnetic material.
In order to study the superparamagnetic state of a single particle, it is simply
necessary to measure the particle's magnetization as a function of time. This
is called telegraph noise measurement as stochastic fluctuations between two
states are expected. According to the Neel-Brown model, the mean time T
spent in one state of magnetization is given by an Arrhenius law of the form of
Eq. (3. 20). As T increases exponentially with decreasing temperature, it is
very unlikely that an escape process will be observed at low temperature.
However, applying a constant field in direction of a hard axis (hard plane) of
magnetization reduces the height of the energy barrier (Fig. 3.22). When the
energy barrier is sufficiently small, the particle's magnetization can fluctuate
between two orientations which are close to a hard axis (hard plane) of
106 Wolfgang Wernsdorfer

magnetization. The time spent in each state follows an exponential switching


probability law as given by Eqs. (3.19) - (3. 21)
21> with a
ex =2. Note that for a
slightly asymmetric energy potential, one switching probability can be so long
that two-level fluctuation becomes practically unobservable.

3.4.3 Experimental Evidence for the Neel-Brown


eel-Brown Model

The Neel-Brown model is widely used in magnetism, particularly in order to


describe the time dependence of the magnetization of collections of particles,
thin films, and bulk materials. However until recently, all the reported
measurements, performed on individual particles, were not consistent with the
Neel-Brown theory. This disagreement was attributed to the fact that real
samples contain defects, ends and, surfaces that could play an important, if
not dominant, role in the physics of magnetization reversal. It was suggested
that the dynamics of reversal occur via a complex path in configuration space,
and that a new theoretical approach is required to provide a correct
description of thermally activated magnetization reversal even in single-domain
ferromagnetic particles (Ledermann et al., 1994; Wernsdorfer et al.,
1995b). Similar conclusions were drawn from numerical simulations of the
magnetization reversal (R (Richards
ichards et al.,
aI., 1996; Gonzalez et aI., 1996 1996;;
Garcia-Pablos et al. , 1996; Hinzke and Nowak, 1998; Boerner and Bertram,
1997) .
A few years later, micro-SQUID measurements on individual Co
nanoparticles showed for the first time a very good agreement with the Neel-
Brown model by using waiting time, switching field, and telegraph noise
measurements (Wernsdorfer et al. , 1997b; Wernsdorfer et al. , 1997c; Jamet
et al., 2001). It was also found that sample defects, especially sample
oxidation, playa crucial role in the physics of magnetization reversal.
In the following subsections, we review some typical results concerning
nanoparticles (Section 3. 4. 3. 1 ), clusters (Section 3. 4. 3. 2) and wires
(Section 3.4.3.3).
3. 4. 3. 3). In Section 3.4.3.4,
3. 4. 3. 4, we point out the main deviations from
the Neel-Brown model which are due to defects.

3.4.3. 1 Application to Nanoparticles

One of the important predictions of the Neel-Brown model concerns the


exponential not-switching probability P ( t) (Eq. (3. 19 which can be
measured directly via waiting time measurements (Section 3. 4. 2. 1). At a
given temperature, the magnetic field is increased to a waiting field H w which
is close to the switching field. Then, the elapsed time is measured until the
magnetization switches. This process is repeated several hundred times, in
order to obtain a waiting time histogram. The integral of this histogram gives
the not-switching probability P (t) which is measured at several temperatures
T and waiting fields H w . The inset of Fig. 3.23
3. 23 displays typical measurements
Classical and Quantum Magnetization Reversal ..
... 107

of P (t) performed on a Co nanoparticle. All measurements show that P (t) is


given by an exponential function described by a single relaxation time T.
T.
143.0

142.8

p
1='
5~ 142.6

~
142.4

142.2'-----_----':-_---:.L-_~--~-----,L,_____-----...J
142.2'-----_----':-_ _,-L:-_--:-':,.--_~---:-'=__---"
o 5 10 25 30

Figure 3. 23 Scaling plot of the mean switching time T( T, H w ) for several waiting fields
H w and temperatures(O.
temperatures( O. 1 s< T ( T, H w ) <60 s)for a Co nanoparticle. The scaling yields
1O- 9 s. Inset: Examples of the probability of not-switching of magnetization as a
To ""'3 x 10-
To::::::3
function of time for different applied fields and at 0.5
O. 5 K. Full lines are data fits with an
exponential function: P ( t) = e-
e - II'/ T
T

The validity of Eqs. <3.5)


(3.5) and (3.20) is tested by plotting the waiting
field H w as a function of [T In (T / To) J2I3 ]2/3 (a = 3/2 because the field was
( ex =
appl ied at about 20 from the easy axis of magnetization (Eq. (3.
applied (3.5. 5)). If the
Neel-Brown model applies, all points should collapse onto one straight line
(master curve) by choosing the proper values for To. To Figure 3.23
3. 23 shows that
T( T, H w ) falls on a master curve provided that To
the data set T( ""::::3 x 10- 9 s.
To::::::::3
The slope and intercept yield the values Eo =214,000 K and H~w = 143. 143.05 05 mT. The
energy barrier Eo can be approximately converted to a thermally "activated
volume" by using V""::::Eo/(JJoMsH~w)""::::(25
V::::::::Eo/(J.1oMsH~w)::::::::(25 nm)3 which is very close to the
particle volume estimated by SEM. This agreement is an important
confirmation of a magnetization reversal by uniform rotation. The result of the
waiting time measurements are confirmed by switching field and telegraph
noise measurements (Wernsdorfer et al., 1997b, 1997c). The field and
temperature dependence of the exponential prefactor To is taken into account
in Coffey et al. (1998).
3.4.3. 2 Application to Co Clusters
Figure 3. 24 presents the angular dependence of the switching field of a 3 nm
Co cluster measured at different temperatures. At O. 03 K, the measurement is
very close to the standard Stoner-Wohlfarth astroid (Fig. 3.3).
3. 3). For higher
temperatures the switching field becomes smaller and smaller. It reaches the
origin at about 14 K, yielding the blocking temperature T B = 14 K of the cluster
magnetization. T B is defined as the temperature for which the waiting time Atb. t
108 Wolfgang Wernsdorfer

becomes equal to the relaxation time T of the particle's magnetization at H = = O.


TB B can be used to estimate the total number N iol tot of magnetic Co atoms in the
cluster. Using an Arrhenius-like law (see Eq. (3. 20 which can be written as
(K at Not I k BT B) , where To lI is the attempt frequency typically
I:,. t = T = To exp (KaINollkBTB)'
t.t=T=To
between 10 10
1010 Hz and 1011 Hz (Respaud et al., 1998), K alat is an effective
anisotropy energy per atom and k B B is the Boltzmann constant. Using the
expression of the switching field at T = = 0 K and for == 0 : !Jo e = 2K all
sw =
J.10 H sw !Jat =
atl J.1al =
0.3T
0.3 T (Fig.3.24),
(Fig.3. 24), the atomic moment !Jat = 1. 7!JB'
J.1al= t.t=O.Ol
7J.1B' I:,.t=O. s, To = 10-los,
01 s,To=10-los,
and T BB= =
= 14 K, we deduce N 101tot""'" 1500, which corresponds very well to a 3 nm
Co cluster (Fig. 3.6).
0.3

0.04K
Ts =14K
0.2

0.1
12K

i='
E="
~
~

::t'
0:;'
~
::

--0.1
--D. I

--0.2
--D. 2

--0.3
--D.3
--0.3
--D.3 --0.2
--D.2 --0.1
-0.1 0 0.1 0.2 0.3
PoHy (T)

Figure 3. 24 Temperature dependence of the switching field of a 3 nm Co cluster,


measured in the plane defined by the easy and medium hard axes ( Hy-H z plane in
Fig. 3.8).
3. 8). The data were recorded using the blind mode method with a waiting time of the
M = O. 1 s. The scattering of the data is due to stochastic and in good
applied field of 1:11
agreement with Eq. (3. 23).

Figure 3. 25 presents a detailed measurement of the temperature dependence


of the switching field at O and 45 and for three waiting times I:,. t. t. This
measurement allows us to check the predictions of the field dependence of the
barrier height. Eq. (3. 22) predicts that the mean switching field should be
proportional to T 1/a
l/a
where a
ex depends on the direction of the applied field
(Eqs. (3.5) - (3.10:
(3. 10: a = e
= 2 for == O and 90, and a == 3/2 for all other
angles which are not too close to = e
= O and 90. We found agreement with this
model (Fig. 3.26).
Classical and Quantum Magnetization Reversal.
Reversal . . . 109

0.16

o 5
T(K)
10 15

Figure 3. 25 Temperature dependence of the switching field of a 3 nm Co cluster,


measured at 0' and 45'. The data were recorded using the blind mode method with
different waiting time M of the applied field. The scattering of the data is due to
stochastics and is in agreement with Eq. (3. 23).

0.04

o 0.2 0.4 0.6


(T/T
(T/TB)l/a
B )lIa

Figure 3. 26 Temperature dependence of the switching field of a 3 nm Co cluster as in


Fig. 3.25 but plotted as a function of (T / T BB)) l/a with T B = 14 K and a = 2 or 3/2 for
e= 0' and 45', respectively, and for three waiting times .tJ.t:..t.t.

3.4.3.3
3.4. 3. 3 Application to Ni Wires

Electrodeposited wires (with diameters ranging from 40 nm to 100 nm and


lengths up to 5000 nm) were studied (Wernsdorfer et ai., aI., 1996a, 1997a)
using the micro-SQUID technique (Section 3. 3. 2. 1). For diameter values
under 50 nm, the switching probability as a function of time could be described
by a single exponential function (Eq. (3. 19. The mean waiting time T
followed an Arrhenius law (Eq. (3. 20 as proposed by the Neel-Brown
model. Temperature and field sweeping rate dependence of the mean
switching field could be described by the model of Kurkijarvi
110 Wolfgang Wernsdorfer

3. 4. 2. 2) which is based on thermally assisted magnetization


(Section 3.4.2.2)
reversal over a simple potential barrier. These measurements allowed us to
estimate an activation volume which was two orders of magnitude smaller than
the wire volume. This confirmed the idea of the reversal of the magnetization
caused by a nucleation of a reversed fraction of the cyl inder, rapidly
propagating along the whole sample. This result was also in agreement with a
micromagnetic model of Braun (1999).
A pinning of the propagation of the magnetization reversal occurred for a
few samples, where several jumps were observed in the hysteresis curves.
The pinning of a domain wall was probably due to structure defects. The
dynamic reversal properties of depinning were quite different from those of
nucleation of a domain wall. For example, the probabil ity of depinning as a
probability
function of time did not follow a single exponential law. A similar effect was
also observed in single submicron Co particles having one domain wall
(Wernsdorfer et a!., aI., 1996b), showing a domain wall annihilation process
(Section 3. 3. 2. 3).

3.4.3.4 Deviations from the Neel-Brown Model


Anomalous magnetic properties of oxidized or ferrimagnetic nanoparticles have
been reported previously by several authors (Berkowitz et al., 1975;
Richardson et al., 1991). These properties are, for example, the lack of
saturation in high fields and shifted hysteresis loops after cooling in the
presence of a magnetic field. These behaviors have been attributed to
uncompensated surface spins of the particles and surface spin disorder
(Kodama et aI.,
a!., 1996; Kodama, 1999).
Concerning our single-particle studies, we systematically observed aging
effects which we attribute to an oxidation of the surface of the sample, forming
antiferromagnetic CoO or NiO (Wernsdorfer et al., al. , 1995b, 1997a). We found
that the antiferromagnetic coupling between the core of the particle or wire and
its oxidized surface changed the dynamic reversal properties. For instance,
we repeated the measurements of the magnetization reversal of a Ni wire two
days after fabrication, six weeks after fabrication, and finally three months
after fabrication (Wernsdorfer et al. , 1997a). Between these measurements,
the wire stayed in a dry box. The quasi-static micro-SQUID measurements did
reveal only small changes. The saturation magnetization measured after six
weeks was unchanged and was reduced by one to two percent after three
months. The angular dependence of the switching field changed only slightly
also. The dynam
dynamicic measurements showed the aging effects more clearly, as
evidenced by
(1) A nonexponential probability of not switching;
(2) An increase of the width of the switching field distributions;
(3) A decrease of the activation energy.
We measured a similar behavior on lithographic fabricated Co particles
with an oxidized border (Wernsdorfer et al.,
al. , 1995b).
Classical and Quantum Magnetization Reversal ...
.. . 111

Figure 3. 27 presents the angular dependence of switching fields of a 3 nm


Fe cluster having a slightly oxidized surface. Huge variation of the switching
fields can be observed, which might be due to exchange bias of frustrated spin
configurations at the surface of the cluster (Fig. 3.28).
0.20

0.15
E
i~ 0.10

0.05':-:--"--'---::-'-::-::-_ _----',-_ _-----,,...,.,:--_--7


0.05':-:--"-"'----=-'":c::----7------:c"=-=----7
-0.1 0 0.05 0.1
J1 oHyy (T)
J10H

Figure 3. 27 Angular dependence of switching fields of a 3 nm Fe cluster having


(probably) a slightly oxidized surface. Each point corresponds to one of the 10,000
switching field measurements. The huge variations of the switching field might be due to
exchange bias of frustrated spin configurations. However, quantum effects like those
described in Section 3. 5. 4 are not completely excluded.

~ 019~~~;
E ~,~~;,: ,wi~,";~
C ::: :~ ;'~i'ld' '~
:r: 0.18
~
,
. switching fields
'~"'l'I--_---='_---
....".,~----=------

:: Negative
switching fields

0.17

-0.01 -0.005 0.005 0.01

Figure 3.28 Details of the angular dependence of switching fields of a 3 nm Fe cluster


having (probably) a slightly oxidized surface. Each point corresponds to one of the
3000 switching field measurements. Stochastic fluctuation between different switching
field distributions are observed. The "mean" hysteresis loop is shifted to negative
fields.

We propose that the magnetization reversal of a ferromagnetic particle


with an antiferromagnetic surface layer is mainly governed by two mechanisms
which are both due to spin frustration at the interface between the
ferromagnetic core and the antiferromagnetic surface layer(s). The first
112 Wolfgang Wernsdorfer

mechanism may come from the spin frustration differing slightly from one cycle
to another, thus producing a varying energy landscape. These energy
variations are less important at high temperatures when the thermal energy
(k B T) is much larger. However, at lower temperatures the magnetization
reversal becomes sensitive to the energy variations. During the hysteresis
loop the system randomly chooses a path through the energy landscape which
leads to broad switching field distributions. A second mechanism may become
dominant at high temperatures: the magnetization reversal may be governed
by a relaxation of the spin frustration, hence by a relaxation of the energy
barrier. This relaxation is thermally activated-that is, slower at lower
temperatures.

3.5 Magnetization Reversal by Quantum Tunneling

Studying the boundary between classical and quantum physics has become a
very attractive field of research which is known as "mesoscopic" physics
(Fig. 3.1).
3. 1 ). New and fascinating mesoscopic effects can occur when
characteristic system dimensions are smaller than the length over which the
quantum wave function of a physical quantity remains sensitive to phase
changes. Quantum interference effects in mesoscopic systems have, until
now, involved phase interference between paths of particles moving in real
space as in SQUIDs or mesoscopic rings. For magnetic systems, similar
effects have been proposed for spins moving in spin space, such as CD
magnetization tunneling out of a metastable potential well or CV ~ coherent
tunneling between classically degenerate directions of magnetization (Leggett
et al., 1987; Gunther and Barbara, 1995).
We have seen in the previous sections that the intrinsic quantum character
of the magnetic moment can be neglected for nanoparticles with dimensions of
the order of the domain wall width 5 {5 and the exchange length A-that is,
particles with a collective spin of S = 10 5 or larger. However, recent
measurements on molecular clusters with a collective spin of S = = 10 suggest
that quantum phenomena might be observed at larger system sizes with
S 1O. Indeed, it has been predicted that macroscopic quantum tunneling
tunnel ing of
magnetization can be observed in magnetic systems with low dissipation. In
this case, it is the tunneling
tunnel ing of the magnetization vector of a single-domain
particle through its anisotropy energy barrier or the tunneling of a domain wall
through its pinning energy. These phenomena have been studied theoretically
and experimentally (Gunther and Barbara, 1995).
The following sections review briefly the most important results
concerning the observed quantum phenomena, which are mesoscopic model
systems in molecular clusters, to test quantum tunnel ing theories and the
Classical and Quantum Magnetization Reversal ... 113

effects of the environmental decoherence. Their understanding requires a


knowledge of many physical phenomena, and they are therefore particularly
interesting for fundamental studies. We then focus on magnetic quantum
tunneling (MQT) studied in individual nanoparticles or nanowires. We
concentrate on the necessary experimental conditions for MQT and review
some experimental results which suggest that quantum effects might even be
= 10 5 or larger.
important in nanoparticles where S =

3.5.1 Quantum Tunneling of Magnetization in Molecular Clusters


Magnetic molecular clusters are the final point in the series of smaller and
smaller units from bulk matter to atoms. Until now, they have been the most
promising candidates for observing quantum phenomena since they have a
well-defined structure with well-characterized spin ground state and magnetic
anisotropy. These molecules are regularly assembled in large crystals where
all molecules often have the same orientation. Hence, macroscopic
measurements can give direct access to single molecule properties. The most
prominent examples are a dodecanuclear mixed-valence manganese-oxo-
(Sessol i et aI., 1993), and an
cluster with acetate ligands, Mn12 acetate (Sessoli
m)
octanuclear iron( ill) oxo-hydroxo cluster of formula [Fe 022(OH)
[FeBBO (OH)12
12 (tacn)6J B+ ,,
(tacn) 6 JB+
FeB (Barra et aI., 1996), where tacn is a macrocyclic ligand (Fig. 3.29).
Both systems have a spin ground state of S = 10, and an Ising-type magneto-
crystalline anisotropy, which stabilizes the spin states where M M= = 10 and
generates an energy barrier for the reversal of the magnetization of about 67 K
for Mn12 acetate and 25 K for FeB'

Figure 3. 29 Schematic view of the magnetic core of the Fea FeB cluster. The oxygen
atoms are black, the nitrogen atoms are gray, and carbon atoms are white. For the
sake of clarity, only few hydrogen atoms are shown as small spheres. The arrows
represent the spin structure of the ground state S = 10 as experimentally determined
through polarized neutron diffraction experiments.
114 Wolfgang Wernsdorfer

Fea is particularly interesting because its magnetic relaxation time


Fes
becomes temperature independent below 360 mK, showing for the first time
that a pure tunneling
tunnel ing mechanism between the only populated M = = 10 states
is responsible for the relaxation of the magnetization (Sangregorio et al. ,
1997) .
The simplest model describing the spin system of Fes
Fea molecular clusters
(called the giant spin model) has the following Hamiltonian:

H =- DS~ + E(S;
E(S~ - S~)SP + 9lJslJoS H (3.25)
Where S x' S y and S z are the three components of the spin operator, 0 and
E are the anisotropy constants, and the last term of the Hamiltonian describes
the Zeeman energy associated with an applied field H. This Hamiltonian
defines hard, medium, and easy axes of magnetization in x, y and z
directions, respectively. It has an energy level spectrum with 2 S + 1 = 21
values which, to a first approximation, can be labeled by the quantum numbers
M= -10,
- 10, -9, "', 10. The energy spectrum, shown in Fig. 3. 30, can be
obtained by using standard diagonalization techniques of the (21 x 21) matrix
describing the spin Hamiltonian S = 10. At H = 0, the levels M = 10 have
the lowest energy. When a field H is applied, the energy levels with MO
increase, whi Ie those with M
while 0 decrease. Therefore, different energy values
MO
can cross at certain fields. It turns out that for Fes
Fea the levels cross at fields
givenbYlJoH ~ n x O.22
given by lJo H nn ",,=,n O. 22 T, with n=l,
n = 1, 2, 3 .... The inset of Fig.3.30
Fig. 3. 30
displays the details at a level crossing where transverse terms containing S x or
S y spin operators turn the crossing into an "avoided level crossing. " The spin
S is "in resonance" between two states when the local longitudinal field is

10

01---
----
--

-20

-30 '--"-_.......,,_----''''-_...l.-:'"------'-_....::>...L.-_-'
L--'o.._....L.:>_---L..::>-_-'---''''------'-_....::>...L.-_-'
o 1.5 2 2.5 3
lloH,
110Hz (T)

Figure 3.30 Zeeman diagram of the 21 levels of the S = 10 m!"nifold m~nifold of Fee as a
function of the field applied along the easy axis (Eq. (3. 25) ) . From bottom to top, the
levels are labeled with quantum numbers M M = - 10, - 9, "', 1O. The levels cross at
fields given by 1J0 ~ n x O. 22 T, with n = 1, 2, 3 .... The inset displays the detail at
/.10 H n """
a level crossing where the transverse terms (terms containing S x or/and S y spin
operators) turn the crossing into an avoided level crossing. The greater the gap .c., ..1 , the
higher is the tunnel rate.
Classical and Quantum Magnetization Reversal ..
... 115

close to an avoided level crossing.


The effect of these avoided level crossings can be seen in hysteresis loop
measurements (Fig. 3. 31). When the applied field is near an avoided level
crossing, the magnetization relaxes faster, yielding steps separated by
plateaus. As the temperature is lowered, there is a decrease in the transition
rate due to reduced thermal-assisted tunneling. The hysteresis loops become
temperature independent below O. 35 K demonstrating quantum tunneling at the
lowest energy levels.

1.0 / .... - / .
( //
( I /
0.5
II 0.7 K
I "
-----J--/
':{ I 0.5 Kr--c,./
I I
0o I
~ I /1 K I I,'
/1 i
-{l.5
-0.5 / I I i
/ I ):
.. / I I I
I'~:"I /'" /--
-1.0
-1'-:.2:--~--o""'.--::6---0':------'O:-'-.6"'-------:'1.2
-1.2 -{l.6 0 0.6 1.2
J,1{)Hz (T)
J1oH
Figure 3. 31 Hysteresis loops of a single crystal of Fee
FeB molecular clusterat different
temperatures. The longitudinal field (z-direction) was swept at a constant sweeping
rate of O. 014 Tis. The loops display a series of steps, separated by plateaus. As the
temperature is lowered, there is a decrease in the transition rate due to reduced thermal
assisted tunneling. The hysteresis loops become temperature independent below
0.35 K.
K, demonstrating quantum tunneling at the lowest energy levels.

The energy gap at an avoided level crossing, the so-called "tunnel


splitting" .1,
Ll, can be tuned by an applied transverse field via the S xH xx and
S yH y Zeeman terms. It turns out that a field in the Hx H x direction (hard
anisotropy direction) can periodically change the tunnel splitting .1.
Ll. In a semi-
classical description (Berry, 1984; Haldane, 1988), these oscillations are
due to constructive or destructive interference of quantum spin phases of two
tunnel paths (Garg, 1993). Such interference effects have been observed for
the first time in FeB
Fea clusters (Wernsdorfer and Sessoli,
Sessol i, 1999). Furthermore,
parity effects were observed when comparing the transitions between different
energy levels of the system (Wernsdorfer and Sessoli, 1999) which are
analogous to the parity effects between systems with half integer or integer
al., 1992; von Delft and Hendey, 1992). Hence, molecular
spins (Loss et aI.,
chemistry has had a large impact on research into quantum tunneling of
magnetization on molecular scales.
116 Wolfgang Wernsdorfer

3. S. 2
3.5.2 Quantum Tunneling of Magnetization in Individual
Single-Domain Nanoparticles
The following sections focuses on MOT studied in individual nanoparticles or
nanowires where the complications due to distributions of particle size, shape,
etc. on are avoided. The experimental evidence for MOT in a single-domain
particle or in assemblies of particles is still a controversial subject. We shall
therefore concentrate on the necessary experimental conditions for MOT and
review some experimental results which suggest that quantum effects might
even be important in nanoparticles with S = 10 5 or larger. We start by
reviewing some important predictions concerning MOT in a single-domain
particle.
On the theoretical side, it has been shown that in small magnetic
particles, a large number of spins coupled by strong exchange interaction can
tunnel through the energy barrier created by magnetic anisotropy. It has been
proposed that there is a characteristic crossover temperature T c below which
the escape of the magnetization from a metastable state is dominated by
quantum barrier transitions, rather than by thermal over barrier activation.
Above T c the escape rate is given by thermal over barrier activation
(Section 3. 4).
4) .
In order to compare experiment with theory, predictions of the crossover
temperature T c and the escape rate rI qt in the quantum regime are relevant.
Both variables should be expressed as a function of parameters that can be
changed experimentally. Typical parameters are the number of spins S,
effective anisotropy constants, applied field strength and direction, coupling to
the environment (dissipation), etc. Many theoretical papers have been
publ ished during the last few years (Gunther and Barbara, 1995). We discuss
published
here a result specially adapted for single-particle measurements, which
concerns the field dependence of the crossover temperature T c .
The crossover temperature T c can be defined as the temperature where
the quantum switching rate equals the thermal one. The case of a magnetic
particle, as a function of the applied field direction, has been considered by
several authors (Zaslavskii, 1990; Miguel and Chudnovsky, 1996; Kim and
Hwang, 1997). We have chosen the result for a particle with biaxial
anisotropy, as the effective anisotropy of most particles can be approximately
described by strong uniaxial and weak transverse anisotropy. The result
according to Kim can be written in the following form (Kim and Hwang,
1997) :
1
1/66
I1 cose 1 /
Tc(e) CC
OC IJoH// EE 1I/4
lJoH// /
4
)1
./1 + 0(1 +1 cose 122/3
3
/ )
(1 cos e I122/3
+ I1 cose 3 (3.26)
(1 / )
Classical and Quantum Magnetization Reversal. . . 117

where IJo H IIII = K


J.1 o H II I M ss and IJo
KII/M H1-1- = K
J.1 o H 1- I M s are the parallel and transverse
K1-/Ms
anisotropy fields given in tesla, K IIII and K 1- are the parallel and transverse
anisotropy constants of the biaxial anisotropy, e is the angle between the easy
axis of magnetization and the direction of the applied field, and E = (1 HI/
( 1- H
H~w). Eq. (3. 26) is valid val id for any ratio a = H 1- I/ HII' HII. The proportionality
proportional ity
coefficient of Eq. (3. 26) is of the order of unity (T c is in units of Kelvin) and
depends on the approach used for calculation (K im and Hwang, 1997).
Eq. (3. 26) is plolted
plotted in Fig. 3.32 for several values of the ratio a. It is valid in
the range .!E<e<n/2-
,j E<e<lr/2 - E.
E.
2
,
\
\ '
\""--
. . _--=~.::. ~~'-~':"~'~""""
---..: .... ".
-- a=O '~;J
........... a=0.2 1
--- a=0.5
--- a=0.8

o 15 30 45 60 75 90
Angle en
e CO)
Figure 3. 32 Normalized crossover temperature T c as given by Eq. (3.26) and for
~ / H II
several values of the ratio a = H -l .

The most interesting feature which may be drawn from Eq. (3. 26) is that
the crossover temperature is tunable using the external field strength and
direction (Fig. 3. 32) because the tunneling
3.32) tunnel ing probability
probabil ity is increased by the
transverse component of the appl ied field. Although at high transverse fields,
T c decreases again due to a broadening of the anisotropy barrier. Therefore,
quantum tunneling experiments should always include studies of angular
dependencies. When the effective magnetic anisotropy of the particle is
fitting parameters. MOT
known, MOT theories give clear predictions with no filting
could also be studied as a function of the effective magnetic anisotropy. In
practice, it is well known for single-particle measurements that each particle is
somewhat different. Therefore, the effective magnetic anisotropy has to be
3. 1. 1).
determined for each particle (Section 3. 3.1.
Finally, it is important to note that most of the MOT theories neglect
damping mechanisms. In Section 3.4. 3. 4. 1 we discussed the case of ohmic
damping (Coffey et a!.al. , 1998), which is the simplest form of damping. More
complicated damping mechanisms might play an important role. We expect
more theoretical work on this in future.
118 Wolfgang Wernsdorfer

3. S.
5. 3 Magnetization Measurements of Individual Single-Domain
Nanoparticles and Wires at Very Low Temperatures
In order to avoid the compl
complications
ications due to distributions of particle size, shape,
etc. , some groups have tried to study the temperature and field dependence of
magnetization reversal of individual magnetic particles or wires. Most of the
recent studies were done using magnetic force microscopy at room
temperature. Low-temperature investigations were mainly performed via
3. 2).
resistance measurements (Section 3.2).
The first magnetization measurements of individual single-domain
nanoparticles at low temperature (0. 1 K < T < 6 K) were presented by
Wernsdorfer et al. (1995a). The detector, a Nb microbridge-DC-SOUID,
(see Section 3. 2. 2) and the particles studied (ellipses with axes between 50
and 1000 nm and thickness between 5 and 50 nm) were fabricated using
electron beam lithography. Electrodeposited wires (with diameters ranging
from 40 to 100 nm and lengths up to 5000 nm) were also studied (Wernsdorfer
et aI., 1996a, 1997a). Waiting time and switching field measurements
(Section 3. 4. 2) showed that the magnetization reversal of these particles and
wires results from a single thermally activated domain wall nucleation,
followed by a fast wall propagation reversing the particle's magnetization. For
nanocrystalline Co particles of about 50 nm and below 1 K, a flattening of the
temperature dependence of the mean switching field was observed which could
not be explained by thermal activation. These results were discussed in the
context of MOT. However, the width of the switching field distribution and the
probability of switching are in disagreement with such a model because
nucleation is very sensitive to factors like surface defects, surface oxidation,
and perhaps nuclear spins. The fine structure of pre-reversa'l magnetization
states is then governed by a multivalley energy landscape (in a few cases
distinct magnetization reversal paths were effectively observed (Wernsdorfer
aI., 1995b and the dynamics of reversal occurs via a complex path in
et al.,
configuration space.
Coppinger et al. (1995) used telegraph noise spectroscopy to investigate
two-level fluctuations (TLF) observed in the conductance of a sample
containing self-assembled ErAs quantum wires and dots in a semi-insulating
GaAs matrix. They showed that the TLF could be related to two possible
magnetic states of a ErAs cluster and that the energy difference between the
two states was a linear function of the magnetic field. They deduced that the
ErAs cluster should contain a few tens of Er atoms. At temperatures between
O. 35 and 1 K, the associated switching rates of the TLF were thermally
activated, whilst below O. 35 K the switching rate becames temperature-
independent. Tunneling of the magnetization was proposed in order to explain
the observed behavior.
Classical and Quantum Magnetization Reversal. . . 119

Some open questions remain: What is the object that is really probed by
TLF? If this is a single ErAs particle.
particle, as assumed by the authors,
authors. the sWitching
switching
probability should be an exponential function of time. The pre-exponential
factor To l1 (called attempt frequency) was found to lie between 10 3 S-1 s -1 and
106 s- 1 whereas expected values are between 109 s- 1 and 10 12 S - 1. Why must
one apply fields of about 2 T in order to measure two-level fluctuations which
should be expected near the zero field? What is the influence of the
measurement technique on the sample?
By measuring the electrical resistance of isolated Ni wires with diameters
between 20 nm and 40 nm. nm, Hong and Giordano studied the motion of magnetic
domain walls (Hong and Giordano.
Giordano, 1995). Because of surface roughness and
oxidation, the domain walls of a single wire are trapped at pinning centers.
oxidation.
The pinning barrier decreases with an increase in the magnetic field. When the
barrier is sufficiently small.
small, thermally activated escape of the wall occurs.
This is a stochastic process that can be characterized by a switching
(depinning) field distribution. A flattening of the temperature dependence of
the mean switching field and a saturation of the width of the switching field
distribution (rms. deviation C1)CT) were observed below about 5 K. The authors
proposed that a domain wall escapes from its pinning site by thermal activation
at high temperatures and by quantum tunneling To~5
tunnel ing below To ""='5 K.
These measurements pose several questions: What is the origin of the
pinning center which may be related to surface roughness, roughness. impurities,
impurities.
oxidation. etc. ? The sweeping rate dependence of the depinning field,
oxidation, field. as well
as the depinning probability.
probability, could not be measured even in the thermally
activated regime. Therefore.
Therefore, it was not possible to check the validity
val idity of the
Neel-Brown model or to compare measured and predicted width of the
switching field distribution (Eq. (3. 23.
23)). Finally.
Finally, a crossover temperature To
of about 5 K is three orders of magnitude higher than To predicted by current
theories.
Later, Wernsdorfer et al. published results obtained on nanoparticles
Later.
discharge. with dimensions between 10 nm and 30 nm
synthesized by arc discharge,
(Wernsdorfer et al. , 1997b). These particles were single crystall ine, ine. and the
surface roughness was about two atomic layers. Their measurements showed
for the first time that the magnetization reversal of a ferromagnetic
nanoparticle of good quality can be described by thermal activation over a
single-energy barrier as proposed by Neel and Brown (see Section 3. 4. 2) .
The activation volume.
volume, which is the volume of magnetization overcoming the
barrier, was very close to the particle volume.
barrier. volume, predicted for magnetization
reversal by uniform rotation. No quantum effects were found down to O. 2 K.
This was not surprising because the predicted crossover temperatures are
below O. 1 K. The results of Wernsdorfer et al. constitute. the preconditions for
the experimental observation of MQT MOT of magnetization on a single-particle.
Just as the results obtained with Co nanoparticles (Wernsdorfer et al al..,
1997b).
1997b), a quantitative agreement with the Neel-Brown model of magnetization
120 Wolfgang Wernsdorfer

reversal was found on BaFe ,12-2x ,9 nanoparticles (0 < x < 1 )


x 0 19
2-2x Cox Ti x

(Wernsdorfer et al.,
ai., 1997c), which we will call BaFeO, in the size range of
10- 20 nm. However, strong deviations from this model were evidenced for
the smallest particles containing about 105 J.ls J.1s and for temperatures below
O. 4 K. These deviations are in agreement with the theory of macroscopic
quantum tunnel ing of magnetization. Indeed, the measured angular
dependence of T cc ( e) is in excellent agreement with the prediction given by
Eq. O.
(3. 26) (Fig. 3.33). The normalization value T c (45) = = 0.31 K compares
well with the theoretical value of about O. 2 K.
1.2

1.0 '",\
\
~ 0.8
-- a=O
.........
pp a=0.05
--- a=O.l
a=O.1
Particle BaFeO
10nm

15 30 45 60 75 90
Angle en
Figure 3. 33 Angular dependence of the crossover temperature T, T c for a 10 nm
BaFelO.4CoosTiosO,9 S""" 10 5 . The lines are given by Eq. (3.26) for
BaFe\o ,Coo 8Tio 80,9 particle with S"""10
different values of the ratio a = H ~ // H II.
H 1. II The experimental data are normalized by
T c (4S)=0.
T,(45) 31 K.
=0.31 K

Although the above measurements are in agreement with MOT theory, we


should not forget that MOT is based on several strong assumptions. Among
them, there is the assumption of a giant spin; that is, all magnetic moments in
the particle are rigidly coupled together by strong exchange interaction. This
approximation might be good in the temperature range where thermal
activation is dominant, but is it not yet clear if this can be made for very low
energy barriers (see, for example, Section 3. 4. 3. 4). Future measurements
might give us the answer.

3. S. 4
3.5.4 Quantization of the Magnetization
In order to give a definite proof that MOT can occur in a magnetic nanoparticle
we propose to surge for the energy level quantization of its collective spin
state. This was recently evidenced in molecular cluster like FeB having a
collective spin state S = 10 (Section 3. 5. 1). In the case of the BaFeO
=
particles with S ""'" 10 5 (Wernsdorfer et al., 1997c), the field separation
Classical and Quantum Magnetization Reversal...
Reversal. . . 121

IlH ~ H
associated with level quantization is rather small: llH:::>::;: Ha/2S~
a /2S:::>::;: 0.002 mT
where H a is the anisotropy field. However, for a 3 nm Co cluster where S ~
S:::>::;:
3
10 the field separation IlH ~ Ha/2S~O.
llH:::>::;:H a /2S:::>::;:0. 2 mT might be large enough to be
measurable.
Figure 3. 34 displays schematically the field values of resonances between
quantum states of S. When the applied field is ramped in a certain direction,

the resonance might occur for fields Hres~n


H res:::>::;: n X
x ~S co~e'
~s with n = 1, 2, 3, ... e
is the angle between the applied field and the easy axis of magnetization. For
large spins S, tunneling might be observable only for fields which are close to
the classical switching field (Fig. 3.34).
3. 34). The resonance fields could be
evidenced by measuring switching field distributions (inset of Fig. 3. 34) as a
function of the angle e. e.

.-:" .
H
:r:~
NT
0
hx

3. 34 Schematic view of the resonance fields of a giant spin S. The continuous


Figure 3.34
line is the classical switching fields of Stoner-Wohlfarth (Section 3. 3. 1). The inset

llH' ~ ~s
presents schematically a switching field histogram with fJ.H' ~~ CO~8
co~e'.

~800.
Such a study is presented in Fig. 3. 35 for a 3 nm Fe cluster with S :::>::;:800.
IlH ~ H aa /2S is about O. 1 mT whereas the
The estimated field separation llH:::>::;:
width of the switching field distribution is about ten times larger. We observed
sometimes a small periodic fine structure which is close to the expected llH. IlH.
However, this fine structure always disappeared when averaging over more
measurements. A possible reason might be hyperfine couplings that broaden
the energy levels leading to a complete overlap of adjacent energy levels. It is
also important to mention that the switching field distributions were
temperature-independent for O. 04< T <0. 2 K.
In some cases, we observed huge variations of the switching field (Figs. 3.27
and 3. 28) which might be dJedue to exchange bias of frustrated spin configurations.
However, quantum effects are not completely excluded. Future measurements should
focus on the level quantization of collective spin states of S = 102 =
122 Wolfgang Wernsdorfer

Figure 3.35 Angular dependence of the switching field of a 3 nm Fe cluster Where S


"",aDO. The inset presents detailed measurements near the easy axis of magnetization.
"""800.

3.
3.66 Summaries and Conclusions

Nanometer-sized magnetic particles have generated continuous interest as the


study of their properties has proven to be scientifically and technologically
very challenging. In the last few years, new fabrication techniques have led to
the possibility of making small objects with the required structural and
chemical qualities. In order to study these objects, new local measuring
techniques such as magnetic force microscopy, magnetometry based on micro-
Hall probes or micro-SQUIDs, were developed. This led to a new
understanding of the magnetic behavior of nanoparticles.
In this article we reviewed the most important theories and experimental
results concerning the magnetization reversal of single-domain particles and
clusters. Special emphasis is was put single-particle measurements avoiding
complications due to distributions of particle size, shape, etc. Measurements
on particle assembl
assemblies
ies have been reviewed in Dormann et al.,
aI., (1997). We
mainly discussed the low-temperature regime in order to avoid spin
excitations.
Section 3.2 reviewed briefly the commonly used measuring techniques.
Among them, electrical transport measurements, Hall probes, and micro-
SQUID techniques seem to be the most convenient techniques for low-
temperature measurements.
Section 3. 3 discussed the mechanisms of magnetization reversal in single-
domain particles at zero Kelvin. For extremely small particles, the
magnetization should be reversed by uniform rotation of magnetization
Classical and Quantum Magnetization Reversal. . . 123

(Section. 3.3.
3. 3. 1). For somewhat larger particles. a nonuniform reversal mode
occurs like the curling mode (Section 3.3.2.1).
3.3.2. 1). For even larger particles.
particles,
magnetization reversal occurs via a domain wall nucleation process starting in
a rather small volume of the particle (Section 3.3.2.3).
3. 3. 2. 3).
The influence of temperature on the magnetization reversal was reported
in Section 3. 4. We discussed in detail the Neel.
Neel, Brown and Coffey's theory of
magnetization reversal by thermal activation.
Finally, Section 3. 5 showed that for very small systems or very low
Finally.
temperature.
temperature, magnetization can reverse via quantum tunneling. The boundary
between classical and quantum physics has become a very attractive field of
research. This section discussed briefly detailed measurements which
demonstrated that molecular magnets offer an unique opportunity to explore the
quantum dynamics of a large but finite spin. The discussion is focused on the
Fes
FeB molecular magnet were S = 10 because it is the first system where studies
in the pure quantum regime were possible. We showed that the understanding
of the environmental decoherence is one of the most important issues, issues. in
particular for future applications of quantum devices. We then discussed
tunneling in nanoparticles and showed how one might give a definite proof of
their quantum character at low temperature.
In conclusion.
conclusion, the understanding of the magnetization reversal in
nanostructures requires the knowledge of many physical phenomena.
phenomena, and
nanostructures are therefore particularly interesting for the development of new
fundamental theories of magnetism and in modeling new magnetic materials for
permanent magnets or high density recording. Using the quantum character of
nanostructures for applications like" Quantum Computers" will be one of the
major concerns of the next decades.

References
Aharoni. A. IEEE Trans. Mag. 22: 478( 1986)
Aharoni,
Aharoni.
Aharoni, A. An Introduction to the Theory of Ferromagnetism. Oxford
University Press. London (1996)
Aharoni. 1281C 1997)
Aharoni, A. J. Appl. Phys. 82: 1281(
Aharoni, A. J. Appl. Phys. 86: 1042 ((1999)
Aharoni. 1999)
Aharoni. A. J. Appl. Phys. 87: 5526
5526(2000)
(2000)
Apsel. E. , J. W. Emmert,
Apsel, S. E. Emmert. J. Deng,
Deng. and L. A. Bloomfield. Phys. Rev.
Lett. 76: 1441(1996)
1441 (1996)
Bansmann, J.,
Bansmann. Senz, L. Lu.
J . V. Senz. Lu, A. Bettac.
Bettac, and K. Meiweis-Broer. J.
Electron Spectrosc. Relat. Phenom. in press( 1999)
Bardou, N.,
Bardou. N. B. Bartenlian,
Bartenlian. C. Chappert, Megy, P. Veillet.
Chappert. R. Megy. Veillet, J. P.
Renard, F. Rousseaux, M. F. Ravet, J. P. Jamet, and P. Meyer, J.
Renard.
Appl. Phys. 79: 5848( 1996)
Barra, A. -L. , P. Debrunner, D. Gatteschi, C. E. Schulz, and R. Sessol i .
EuroPhys. Lett. 35: 133 ( 1996)
124 Wolfgang Wernsdorfer

Bean, C. P. J. Appl. Phys. 26: 1381(1955)


1381C 1955)
Bean, C. P. and J. D. Livingstone, J. Appl. Phys. 30: 120S(1959)
Berkowitz, A. E., J. A. LahuL Lahut, I. S. Jacobs, L. M. Levinson, and D.
Forester. Phys. Rev. Lett. 34: 594 (1975)( 1975)
Berry, M. V. Proc. R. Soc. London A, 392: 45( 45(1984)
1984)
Billas, I. M. L., A. Chatelain, and W. A. de Heer, J. Magn. Magn. Mat.
168: 64(1997)
64( 1997)
Boerner, E. D. and H. N. Bertram. IEEE Trans. Mag., 33: 3052 (1997)
Bonet, E. , Wernsdorfer, W. , Barbara, B. , Hasselbach, K. , Benoit, A. and
D. Mailly. IEEE Trans. Mag. 34: 979(1998)
Bonet, E., W. Wernsdorfer, B. Barbara, A. Benoit, D. Mailly, and A.
Thiaville. Phys. Rev. Lett. 83: 4188(1999)
Braun, H.-B.
H. -B. J. Appl. Phys. 85: 6172(1999)
6172 ( 1999)
Brown, W. F. J. Appl. Phys. 30: 130S(1959)
Brown, W. F. J. Appl. Phys. 34: 1319 1319(1963a)
( 1963a)
Brown, W. F. Phys. Rev. 130: 1677(1963b)
Bryant, P. and H. Suhl. Appl. Phys. Lett. 54: 78(1989) 78 (1989)
Chang, C. -R. J. Appl. Phys. 69: 2431( 2431C 1991)
Chang, T. and J. G. Chu. J. Appl. Phys. 75: 5553(1994)
Chuang, D. S., 0' Handley, Phys. Rev. B 49:
S. , C. A. Ballentine, and R. C O'Handley,
15,084(1994)
Coffey, W. T. , D. S. F. Crothers, J. L. Dormann, Y. P. Kalmykov and J.
T. Waldron, Phys. Rev. B 52: S2: 15,951(1995)
15 ,951( 1995)
Coffey, W. T. , D. S. F. Crothers, J. L. Dormann, Y. P. Kalmykov, E. C.
Kennedy and W. Wernsdorfer. Phys. Rev. Lett. 80: 5655(1998) 5655 ( 1998)
Coppinger, F. , J. Genoe, D. K. Maude, U. Genner, J. C. Portal, K. E.
Singer, P. Rutter, T. Taskin, A. R. Peaker and A. C. Wright. Phys.
Rev. Lett. 75: 3513(1995)
Dormann, J. L., D. Fiorani, and E. Trone. Adv. Chem. Phys. 98: 283
(1997) .
(1997).
Doudin, B. and J. -P. Ansermet. NanoStructured Mater. 6: 521 521C( 1995)
Lee, C. H. , Hui He, F. J. Lamelas, W. Vavra, C. Uher, and Roy Clarke.
Phys. Rev. B42: 1066 (1990)
Fernandez, A. , M M. Gibbons, M. Wall, and C. Cerjan. J. Magn. Magn. Mat.
190: 71(1998)
71( 1998)
Ferre, R. , K. Ounadjela, J. M. George, L. Piraux, and S. Dubois, Phys.
Rev. B 56: 14,066(1997)
Freemann, A. J. and R. Wu, J. Magn. Magn. Mat. 100: 497(1991)
Frei, H. , S. Shtrikman and D. Treves. Phys. Rev. 106: 446(1957)
Fruchart, O. , J. -P.-Po Nozieres, W. Wernsdorfer, and D. Givord. Phys. Rev.
Lett. 82: 1305(
1305 ( 1999)
Gallagher, W. J., J. , S. S. P. Parkin, Y. Lu, X. P. Bian, A. Marley, K. P.
Roche, R. A. Altman, S. A. Rishton, C. Jahnes, T. M. Shaw, and G.
Xiao. J. Appl. Phys. 81: 3741(1997)
3741 (1997)
Classical and Quantum Magnetization Reversal.
Reversal . . . 125

Garcia-Pablos, D. , P. Garcia-Mochales and N. Garcia. J. Appl. Phys. 79:


6021(1996)
602l< 1996)
205 ( 1993)
Garg, A. EuroPhys. Lett. 22: 205(1993)
Garg, A. Phys. Rev. B 51: 15,592(1995)
Gonzalez, J. M., R. Ramirez, R. Smirnov-Rueda and J. Gonzalez. J. Appl.
Phys. 79: 6479(1996)
6479( 1996)
Gros, V. , S. -F. Lee, G. Faini, A. Cornette, A. Hamzic and A. Fert, J.
Magn. Magn. Mat. 165: 512 512((1997)
1997)
Gueron, S. , M. M. Deshmukh, E. B. Myers and D. C. Ralph, Phys. Rev.
Lett. 83: 4148( 1999)
4148(1999)
Guerret-Piecourt, C. , Y. L. Bouar, A. Loiseau and H. Pascard. Nature 372:
761C 1994)
76l<
Gunther, L. and B. Barbara. Quantum Tunneling of Magnetization-QTM' 94 ,
volume 301 of NA TO ASI Series E: Applied Sciences. Kluwer Academic
Publishers, London( 1995)
Haldane, F. Phys. Rev. Lett. 61: 1029 1029(1988)
( 1988)
Hefferman, S. J. J.,, J. N. Chapman, and S. McVitie, J. Magn. Magn. Mat.
95: 76( 1991)
Hinzke, D. and U. Nowak. Phys. Rev. B 58: 265(1998)
Hong, K. and N. Giordano, J. Magn. Magn. Mat. 151: 396(1995) 396 ( 1995)
Hubert, A. and R. Schafer. Magnetic Domains: The Analysis of Magnetic
Microstructures. Springer-Verlag, Berlin, Heidelberg, New Yorke 1998)
York(1998)
Ishii, Y. J. Appl. Phys. 70: 3765(1991)
Jamet, M. , V. Dupuis, P. Melinon, G. Guiraud, A. Perez, W. Wernsdorfer,
A. Traverse and B. Baguenard. Phys. Rev. B 62: 493 (2000)
493(2000)
Jamet, M. , W. Wernsdorfer, C. Thirion, D. Mailly, V. Dupuis, P. Melinon
and A. Perez Phys. Rev. Lett. 86: 4676 (2001) (200 1)
A. , S. von Molnar, S. Gider and D. Awschalom J. Appl. Phys. 76:
Kent, A.,
6656(1994)
Kim, G.-H. and D. S. Hwang. Phys. Rev. B 55: 8918-8927(1997)
Knowles, J. E. IEEE Trans. Mag. MAG-14: 858(1978)
Kodama, R. H., A. E. Berkowitz, E. J. McNiff Jr. and S. Foner. Phys.
394 ( 1996)
Rev. Lett. 77: 394(1996)
Kodama, R. J. Magn. Magn. Mat. 200: 359(1999) 359 ( 1999)
Kohl, C. and G. F. Bertsch. Phys. Rev. B60: 4205 ( 1999)
Kurkijarvi, J. Phys. Rev. B6: B 6: 832(1972)
Ledermann, M. , S. Schultz and M. Ozak Ozaki.i. Phys. Rev. Lett. 73: 1986 ((1994)
1994)
Ledermann, M.,M. , R. O'Barr and S. Schultz. IEEE Trans. Mag. 31: 3793 Lee
al. , 1990 (1995)
et a!.
Leggett, A. J. , S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg
and W. Zwerger. Rev. Mod. Phys. 59: 1(1987) l< 1987)
Lok, J., A. Geim, J. Maan, S. Dubonos, L. T. Kuhn and P. Lindelof.
Phys. Rev. B 58: 12 12,201(1998)
,20l< 1998)
Loss, D. , D. P. DiVincenzo and G. Grinstein. Phys. Rev. Lett. 69: 3232
, 26
126 Wolfgang Wernsdorfer

(1992)
Meier, J.,J. , B. Doudin and J. -P.-Po Ansermet. J. Appl. Phys. 79: 6010(1996)
6010 (1996)
Meiweis-Broer, K. Phys. BI. 55: 21(1999)21C 1999)
Miguel, M. C. and E. M. Chudnovsky Phys. Rev. B 54: 388-394(1996) 388 - 394 ( 1996)
Morrish, A. H. and S. P. Yu. Phys. Rev. 102: 670(1956)
Neel, L. C. R. Acad. Science 224: 1550(1947)
Neel, L. Ann. Geophys. 5: 99( 99 ( 1949a)
Neel, L. C. R. Acad. Science 228: 664(1949b)
664 ( 1949b)
Pastor, G. M., J. Dorantes-Davila, S. Pick and H. Dreysse. Phys. Rev.
Lett. 75: 326( 1995)
Perez, A. , P. Melinon, V. Depuis, P. Jensen, B. Prevel, J. Tuaillon, L.
Bardotti, C. Martet, M. Treilleux, M. Pellarin, J. Vaille, B. Palpant and
J. Lerme, J. Phys. D 30: 709( 1997)
709(1997)
Pfeiffer, H. Phys. Status Solidi 118: 295(1990a)
Pfeiffer, H. Phys. Status Solidi 122: 377(1990b)
Respaud, M., J. M. Broto, H. Rakoto, A. R. Fert, L. Thomas and B.
Barbara. Phys. Rev. B 57: 2925 ( 1998)
Richards, H. L. , S. W. Sides, M. A. Novotny and P. A. Rikvold, J. Appl.
Phys. 79: 5749( 1996)
Richardson, J. T. , D. I. Yiagas, B. Turk, J. Forster, and M. V. Twigg J.
Appl. Phys. 70: 6977 ( 1991)
Richter, H. J. J. Appl. Phys. 65: 9(1989)
Ri..ihrig, M. , Bartsch, W. , Vieth, M. , and Hubert, A. (1990). IEEE Trans.
Rlihrig,
Mag. , MAG-26: 2807 .
Salling, C.,
C. , S. Schultz, I. McFadyen and M. Ozaki IEEE Trans. Mag. 27:
5185(1991)
Sangregorio, C. , T. Ohm, C. Paulsen, R. Sessoli and D. Gatteschi. Phys.
Rev. Lett. 78: 4645(1997)
4645 ( 1997)
Schelp, L. L.,, A. Fert, F. Fettar, P. Holody, S. F. Lee, J. L. Maurice, F.
Petroff and A. Vaures Phys. Rev. B 56: R5747. R57 47. (1997)
Schweinbock, T. , D. Weiss, M. Lipinski and K. Eberl. J. Appl. Phys. 87:
6496(2000)
Sessoli, R., D. Gatteschi, A. Caneschi, and M. A. Novak Nature 365:
141- 143( 1993)
Stoner, E. C. and E. P. Wohlfarth, Philos. Trans. London Ser. A 240: 599
(1948) reprinted in IEEE Trans. Magn. MAG-27, 3475 (1991)
Thiaville, A. J. Magn. Magn. Mat. 182: 5( 1~98)
5(19.98)
Thiaville, A. Phys. Rev. B 61: 12,221(2000)
Tonomura, A. , T. Matsuda, J. Endo, T. Arii, and K. Mihama. Phys. Rev.
B 34: 3397
3397(1986)
( 1986)
van den Berg. H. A. M., J. Appl. Phys. 61: 4194( 1987)
4194(1987)
Victora, R. H. Phys. Rev. Lett. 63: 457(1989)
von Delft, J. and C. L. Hendey. Phys. Rev. Lett. 69: 3236 ( 1992)
Wegrowe, J. E. , S. E. Gilbert, D. Kelly, B. Doudin and J. -Po Ansermet.
Classical and Quantum Magnetization Reversal. ..
Reversal ... 127

IEEE Trans. Mag. 34: 903( 903 (1998)


1998)
Wegrowe,
Wegrowe. J. -E.-E ., D. Kelly.
Kelly, A. Franck.
Franck, S. E. Gilbert and J. -Po Ansermet.
Phys. Rev. Lett. 82: 3681(3681C 1999)
Wernsdorfer,
Wernsdorfer. W..W., K. Hasselbach.
Hasselbach, A. Benoit.
Benoit, B. Barbara.
Barbara, D. Mailly.
Mailly, J.
Tuaillon. Perez, V. Dupuis.
Tuaillon, J. P. Perez. Dupuis, J. P. Dupin.
Dupin, G. Guiraud,
Guiraud. and A.
Perez. J. Appl. Phys. 78: 7192 ( 1995a)
Wernsdorfer,
Wernsdorfer. W. W ., K. Hasselbach,
Hasselbach. A. Benoit, Cernicchiaro, D. Mailly.
Benoit. G. Cernicchiaro. Mailly,
B. Barbara and L. Thomas J. Magn. Magn. Mat. 151: 38( 1995b)
Wernsdorfer, W
Wernsdorfer. W.,. K. Hasselbach.
Hasselbach, D. Mailly.
Mailly, B. Barbara.
Barbara, A. Benoit.
Benoit, L.
Thomas and G. Suran. J. Magn. Magn. Mat. 145: 33( 33 <1995c)
1995c)
Wernsdorfer,
Wernsdorfer. W. Magm3tometrie
Magmtometrie d micro-SQUID pour I' etude de particules
ferromagnetiques isolees aux echelles
echeIJes sub-microniques. PhD thesis.
thesis, Joseph
University, Grenoble( 1996)
Fourier University.
Wernsdorfer, W.,
Wernsdorfer. Doudin, D. Mailly.
W . B. Doudin. Mailly, K. Hasselbach.
Hasselbach, A. Benoit.
Benoit, J.
Meier.
Meier, J. -P . Ansermet and B. Barbara. Phys. Rev. Lett. 77: 1873
((1996a)
1996a)
Wernsdorfer, W. , K. Hasselbach, A. Sulpice, A. Benoit, J. -E. Wegrowe,
L. Thomas, B. Barbara and D. Mailly. Phys. Rev. B 53: 3341C1996b) 3341(1996b)
Wernsdorfer, W., K. Hasselbach, A. Benoit, B. Barbara, B. Doudin, J.
Meier, J.-P. Ansermet and D. Mailly. Phys. Rev. B 55: 1155(1997a)
Wernsdorfer, W..W., E. B. Orozco, K. Hasselbach, A. B. B. Barbara, N.
Demoncy, A. Loiseau.
Loiseau, D. Boivin, H. Pascard and D. Mailly. Phys. Rev.
Lett. 78: 1791(1997b)
1791C 1997b)
Wernsdorfer, W. , E. B. Orozco, K. Hasselbach, A. Benoit, D. Mailly, O.
Kubo, H. Nakano and B. Barbara Phys. Rev. Lett. 79: 4014(1997c)
Kubo.
Wernsdorfer W.,
W. , and R. Sessoli.
Sessol i. Science 284: 133(1999)
133 ( 1999)
Zaslavski i, O. B. Phys. Rev. B 42: 992-993(1990)
Zaslavskii, 992 - 993 (1990)

The author is indebted to A. Benoit. E. Bonet Orozco.Orozco, V. Bouchiat,


Bouchiat. I. Chiorescu,
Chiorescu. M.
Faucher. K. Hasselbach,
Faucher, Hasselbach. M. Jamet, Mailly. B. Pannetier,
Jamet. D. Mailly, Pannetier. and C. Thirion for their
experimental contributions and the developement of the ~icro-SQUID
~icro-SQUID technology. The author
acknowledges the collaborations with J. -Ph. Ansermet.
J.-Ph. Ansermet, B. Barbara.
Barbara, A. Caneschi,
Caneschi. A.
Cornia,
Cornia. N. Demoncy,
Demoncy. B. Doudin,
Doudin. V. Dupuis,
Dupuis. O. Fruchart,
Fruchart. D. Galteschi,
Gatteschi. O. Kubo, J. -Po
Kubo. J.-P.
Nozieres. H. Pascard,
Nozieres, Pascard. C. Paulsen, A. Perez, C. Sangregorio, and R. Sessoli. He is also
Coffey. A. Garg, N. Prokof'ev,
indebted to W. T. Coffey, Prokof'ev. P. Stamp, A. Thiaville,
Thiaville. I. Tupitsyn,
and J. Villain for many fruitful and motivating discussions. This work has been supported by
CNRS, DRET, MASSDOTS, and Rhone-Alpes.
4 Micromagnetic Simulation of Dynamic and
Thermal Effects

T. Schrefl, J. Fidler, D. Suess, W. Scholz, V. Tsiantos

4. 1 Introduction

Small magnetic elements are the basic structural units of magneto-electronic


devices (Prinz, 1999) and discrete storage media (Terris et al. , 1999). The
development of magnetic sensors or magnetic memory cells requires a precise
knowledge of the magnetization reversal mechanism of magnetic
nanostructures. Finite element micromagnetics take into account the complex
microstructure of magnetic materials such as edge roughness, grain structure
and particle shape. In combination with magnetic imaging using magnetic force
(Dahlberg and Zhu, 1995) and Lorentz (K (Kirk
irk et ai.,
aI., 1997) microscopy, the
simulations provide a useful tool to characterize the reversal magnetization
reversal process. Both in magnetic recording and in magnetic memory cells as
used in magnetic random access memories, a high data rate is desired.
Numerical micromagnetics can provide a basic understanding of the switching
dynamics. If the particles are sufficiently small, the magnetization reverses by
quasi-uniform rotation (Street and Crew, 1999). Then the energy barrier for
thermally activated switching decreases with decreasing particle volume. Thus
with decreasing size of the structural magnetic units, thermal effects become
important and may influence the switching time.
The time evolution of the magnetization can be computed by solving the
Gilbert equation of motion (Gilbert, 1955). It describes the precession of the
magnetization around the effective magnetic field subject to viscous damping.
The strength of the damping term considerably influences the reversal process
(Kikuchi,
(K ikuchi, 1956). Kikuchi derived the critical value of the damping constant
which minimizes the switching time. The Gilbert damping constant for critical
damping is ex = 1 for spheres, and ex = O. 01 for thin films. In order to include
thermal activation, a random thermal field can be added to the effective field.
The resulting stochastic equation of motion describes the random motion of the
magnetization in thermal equilibrium and eventually across energy barriers.
Section 4. 2 describes the micromagnetic background of the simulations.
Section 4. 3 introduces the basic numerical techniques. Section 4. 4 presents
examples of the switching dynamics and thermal processes in columnar grains,
thin film elements, nanodots, and nanowires.
Micromagnetic Simulation of Dynamic and Thermal Effects 129

4. 2 Micromagnetic Background

4. 2. 1 Equation of Motion

The theoretical treatment of dynamic effects starts from the torque, m x eff ,
X H eff ,

exerted on the magnetic moment m, by the effective field H H eff


eff
This torque will

rotate the magnetic moments of the electrons with respect to the lattice.
According to quantum theory the angular momentum associated with the
magnetic moment m is m/ y, where y is the gyromagnetic ratio of the system,
often close to that of a free electron. The torque equation

:t(;)= = m X H eff (4.1)

describes the motion of the magnetic moment around the effective field.
Equation (4. 1), which describes the gyromagnetic precession of the magnetic
moment, states that the rate of change of the angular momentum with time, t,
equals the torque.
In equilibrium the change of the angular momentum with time and thus the
torque is zero. In order to describe the motion of the magnetic moment
towards equilibrium a viscous damping term can be included. It results from a
(a m/O t), which is added to the effective field. This
dissipative term, - TJ77 (am/a
am/O t, with TJ77
dissipative term is proportional to the generalized velocity, am/a
being a positive constant. The Gilbert equation of motion is

:t(;)=mX(Heff-77~~).
:t(;)= m (H TJ ~~).
X eff - (4.2)

Equations (4. 1) and (4. 2) keep I m I constant. In a more convenient notation,


the Gilbert equation is
am aex m X am
-at ==--II y II m x HHeffeff + TmI
at Iml x at
--m -
at
(4.3)

where the dimensionless Gilbert damping constant a ex = - YTJ


Y77 I m I has been
introduced.
The first term in the right hand side of Eq. (4. 3) describes the
gyromagnetic precession and the second term is the damping term which
describes the motion of the magnetic moment towards the effective field. In
equilibrium the magnetic moment is parallel to the effective field and the
torque, m xXH eff ,
H eff , vanishes. Figure 4. 1 summarizes the basic contributions to

the effective field. It is the sum of the exchange field, the magnetostatic field,
the anisotropy field and the external field. The exchange field and the
130 T. Sehretl et al.

magnetostatic field introduce interactions between neighboring magnetic


moments. At non-zero temperatures a random stochastic field may be
included. The exchange field causes the neighboring magnetic moments to be
aligned parallel to each other, the magnetostatic field breaks large magnetic
particles into smaller magnetic domains, the anisotropy field causes the
magnetic moments to be oriented along certain crystallographic directions,
and the external field rotates the magnetization parallel to its own direction.

\l.!~
o ~11:
... I Thermal activation
...ij.25 0.5
-I
-0.:' U 05 -0,"
I _1- 11.5
1 -I

~ Easy directions
c::>
~ Parallel spins
c::>

~ Fluctuations
c::>
Magnetostatic

~
c::> Domains
~
c::> Rotation

Figure 4. 1 Basic micromagnetic contributions to the effective field.

4. 2. 2 Gibbs Free Energy

In equilibrium the total magnetic Gibbs free energy reaches a local minimum.
The total magnetic Gibbs free energy is the sum of the exhange energy, the
magnetostatic energy, the magnetocrystalline anisotropy energy and the
CBrown, 1963). In a continuum theory, the direction of the
Zeeman energy (Brown,
magnetic moments is described by the magnetization vector M. M is the
magnetic moment (m) Cm) per unit volume. The total energy of a ferromagnetic
particle is a function of the magnetization distribution M (r)
Cr) and the external
field H ext
ext ::

ECM,H
E(M,H ext )) = f[ A2 .:8 (V M;)2 - ~o HdM + f k (M) - 1-10 HextMJd V
Ms 1=1

(4.4)

where M s is the spontaneous magnetization and M; is the i-th component of the


magnetization vector M. A is the exchange constant and f k the magneto-
crystall ine anisotropy energy density. The demagnetizing field, H d
crystalline d ,, can be
Micromagnetic Simulation of Dynamic and Thermal Effects 131

expressed in terms of magnetic volume charges, V M, due to inhomogeneous


magnetization distributions within the magnetic particle and magnetic surface
charges, M n, at grain boundaries and free surfaces with unit surface normal
n. The effective field is the negative variational derivative of Eq. (4. 4) with
respect to the magnetization.

4. 2. 3 Langevin Equation
The effects of thermal motions on a short time scale can be treated numerically
adding a random thermal field to the effective field in Eq. (4. 3). This leads to
the Langevin equation
am
om
at
Jt =-1 y I m x (H eff +H th )
a
+ TmI am
om
at
m x Jt (4.5)

where a is Gilbert damping constant.


The random field, H th , describes the coupling of the magnetic system
with a heat bath. It accounts for the interaction of the magnetic polarization
with the microscopic degrees of freedom which causes the fluctuation of the
magnetization distribution. The fluctuations are assumed to take place on a
much faster time scale than the intrinsic time scale given by the gyromagnetic
ratio and the effective field. The intrinsic time scale as given by the Gilbert
equation of motion follows from the Larmor frequency
w =- yH eff (4.6)
The thermal field is assumed to be a Gaussian random process with the
following statistical properties:
(4.7)

(4.8)
where Hih~; is i-th component of the thermal fluctuation field at magnetic
moment k. The average of the thermal field taken over different realizations
vanishes in each direction i in space. k and I are position indices. The
thermal field is uncorrelated in time and space. The strength of the thermal
fluctuations follow from the fluctuation-dissipation theorem (Brown, 1979):

o = 2ak sB T (4.9)
ylml
Ylml
where k Bs is the Boltzmann constant and T is the temperature.
Figure 4. 2 compares the motion of the magnetization towards equilibrium
for different damping and different temperatures. The figure gives the trace of
the magnetic moment in the x, y plane. The effective field is constant and
parallel to the z-axis of the Cartesian coordinate system. At high damping the
magnetization rotates more directly towards the field direction, as the second
132 T. Schrefl et al.

term in Eq. (4. 3) is dominant. When the precession term becomes dominant,
then the magnetization precesses several times around the field direction
equil ibrium. In the case of low damping and non-zero
before it reaches equilibrium.
temperature the magnetization moves randomly and eventually comes to
thermal equilibrium.

a=1
a=!
/
I\
Hk

,
- -,.
-'f
_M Cobalt
V=4 3 nm 3
~
::;t
:::t y
Minimum x
reversal time
T 300 K
K
M,
a=O.1
a=O.l
(a)

~
a=O.1

::{.
~'" @
Gyromagnetic
:::;t
~'"

Random
walk
precession M
M,x
Mx

(b) (c)

Figure 4. 2 Malian
Motion of the magnetization towards equilibrium. (a) High damping,
(b) low damping, and (c) low damping at non-zero temperature.

4.2.4 Characteristic Length Scales


The competitive effects of the different contributions to the effective field lead
to characteristic length scales on which the magnetization changes its
direction. To minimize the magnetostatic energy magnetic particles may break
up into domains with different directions of the magnetization. Within the
domain wall the magnetization changes its direction on a length which is
comparable with

(4. 10)

where A is exchange constant. In hard magnetic materials with a uniaxial


anisotropy constant K uor with

= ITTT
ON = J = TT/
2A2 =
lJo M s
IT I ex (4.11)
(4. 11)
Micromagnetic Simulation of Dynamic and Thermal Effects 133

in soft magnetic thin films. Here 00 is the Bloch parameter and lex is the
exchange length. The smaller of the two values defines a characteristic length:
The magnetization is often assumed to be uniform within regions smaller than
about one half of the characteristic length. The domain wall width differs from
the characteristic length by a factor of 1T.
"IT.
In numerical micromagnetic simulations the ferromagnetic particle is
subdivided into smaller computational cells. Within each cell the magnetization
is assumed to be uniform. Thus it is possible to associate a rigid magnetic'
moment to each cell. The time evolution of the magnetization follows from the
solution of a coupled system of ordinary equations. Generally one equation of
motion as given by Eq. (4. 3) has to be solved for each computational cell. In
order to resolve the transition of the magnetization between magnetic domains
it is required that the size of the computation cells are smaller than half of the
characteristic length. If this condition is fulfilled, the numerical results are
independent of the grid size (Rave et al. , 1998).

4.3 Numerical Techniques

4. 3. 1 Finite Element Discretization


One possible technique to subdivide the magnetic structure into computational
cells is the finite element method. It has the advantage that complex
microstructures like edge irregularities and polyhedral grains can be modeled
easily (Schrefl, 1999).
1999) . The discretization starts from the continuum
expression for the total energy Eq. (4. 4). The magnetization vector is
interpolated with piecewise linear function on a tetrahedral finite element
mesh. Adaptive refinement methods (Hertel and Kronmi
Kronmiiller,
iller, 1998; Scholz et
aI., 1999) keep the number of elements small while resolving the
magnetization within domain walls. It is possible to assign a magnetic moment
to each node of the finite element grid using a box scheme

m(k) =
= f
Ck )
V(k)
M s (r) dV
Ms(r)dV (4. 12)
V

where the integral has taken over the volume V(k)


V(kJ surrounding node k. The

box volumes fulfill the condition

~V<kJ
~V(k) = fdV' V<k)
V(k) n VV(/) = 0,
W
for k #- I. (4.13)
k

Figure 4. 3a gives an example for an adaptive finite element grid obtained from
the simulation of domain wall motion in a thin specimen of a granular hard
134 T. Sehrefl et al.

magnet (Scholz et al. , 2000). Figure 4. 3b shows the schematics of the box
scheme. The effective field at node k can be approximated using

H(k) =-~ (4. 14)


eff ()m(k)

where E is the total magnetic Gibbs free energy.


With Eqs. (4.4), (4.12), and (4.14) the magnetic moment and the
effective field can be evaluated at the nodes of the finite element mesh. Now
for each node an equation of motion has to be solved. The equations at the
different nodes are coupled by the exchange and magnetostatic field.

(a) (b)

Figure 4.3 Schematics of the space discretization: (a) Tetrahedral finite element mesh
of a granular thin film. The region of fine mesh results from adaptive refinement near a
domain wall. (b) Box volumes and node points.

4.3.22
4.3. Magnetostatic Field Calculation
The exchange field and the magnetostatic field couple the motion of the
magnetic moments at the nodes of the finite element grid. Whereas the
exchange interaction is short range and involves only nearest neighbors, the
magnetostatic interaction is long range. The demagnetizing field, H d , at a
given node depends on the magnetization distribution over the entire mesh. On
regular grids fast fourier transform methods (Ramstock,
(Ramst6ck, 1994) are used to
effectively compute the magnetostatic interactions. The simulation of irregular
grain structure requires unstructured grids. It is possible to eliminate the long
range terms from the equations introducing a magnetic scalar potential, H d =
- V U. The scalar potential follows from the magnetostatic boundary value
problem
V 2U = VM inside the magnet, (4. 15)

V 2U = 0 outside the magnet, and (4. 16)


Micromagnetic Simulation of Dynamic and Thermal Effects 135

(V U inin
(''VU - V U out
'Vu = M n
out)) n = M n.
at the boundary with surface normal n.
(4.17)

Equations (4. 15) to (4. 17) can be solved using a hybrid finite element /
boundary element method (Fredkin and Koehler, 1990). This method is
especially useful for the simulation of the magnetostatic interactions of distinct
magnetic elements, since no mesh is required outside the magnetic particles.
=
The magnetic scalar potential is split into U = U II + U z , where U II accounts for
the divergence of magnetization within the particle and U z is required to meet
the boundary conditions. The latter also carries the magnetostatic interactions
between distinct magnetic particles. U I is zero outside the particle and is the
solution of the Poisson equation within the particle with the boundary
condition, VU,"'lUI n = M n. The potential U z satisfies the Laplace equation
everywhere and shows a jump at the surface of the particle. The computation
of U consists of three steps:
(1) A standard finite element method is used to solve Poisson's equation
for U,
U I .
(2) The potential U z is calculated at the boundary:
= BU 11 , B is an m x m matrix which relates the nodes at the surface to
U z =BU
each other and U I is the vector of the U I values at the surface nodes. The
matrix B is dense and follows from the boundary element discretization of the
double layer operator.
(3) Once U z at the boundary has been calculated, the values of U z within
the particles follow from Laplace's equations with Dirichlet boundary
conditions, which again can be solved by standard finite element technique.

4. 3. 3 Time Integration

After the evaluation of the magnetic moments and the effective fields at the
nodes of the finite element grid a system of coupled ordinary differential
equations have to be solved. In numerical micromagnetic simulations various
time integrations methods have been used. In numerical micromagnetics the
Runge Kutta method or Adams method are suitable for weakly coupled systems
(Mansuripur, 1988; Zhu and Bertram, 1989). Both are successfully used for
the simulation of magnetostatically interacting grains in magnetic recording.
Higher order backward differentiation formula (BDF) methods are most
appropriate for ordinary differential equations resulting from the space
discretization of a partial differential equations. Especially in a highly
exchange coupled system, BDF methods are more efficient than explicit time
integration schemes. BDF methods are implicit and thus require to solve a
nonIinear
nonlinear system of equations at each time step. However, only a few Newton
steps are required to obtain convergence. At each Newton step a linear
system of equations has to be solved. Due to the long-range magnetostatic
interactions, which couple all the nodes of the computational grid, the system
136 T. Sehrefl
Schrefl et al.

matrix of this equation is fully populated. The use of a generalized minimum


residual method (GMRES) avoids the storage of the system matrix. A
considerable speed up for the solution of the linear system can be obtained
with proper preconditioning.
The software package CVaDE
CVODE (Hindmarsh and Petzold, 1995) provides a
general framework to compare different time integration schemes. Figure 4. 4a
gives the CPU time as a function of the simulated time during the magnetization
reversal of a granular thin film element. Figure 4. 4b shows the remanent state
and a snapshot during magnetization reversal. The dark areas are reversed
domains. The comparison of the CPU time for different time integration
schemes confirms that the preconditioned BDF method is faster than the Adams
method. For preconditioning, an approximate Jacobian matrix is passed to
CVaDE which includes the exchange and anisotropy term but omits the
CVODE
magnetostatic contributions. Thus the approximate Jacobian remains sparse
and can be calculated on the fly. Yang and Fredkin (1998) used a similar
approach to simulate magnetization reversal in ellipsoidal particles.
'~~'1"f ,
50,000 j,,!!~~; ,
Adalns
Adams ,I",~:.. ......
',~,~":'" Iv
40 000
40,000 l\'1F 1
I\ J
~ , ";I'!\;,I
~ ';""1]
BDF-Krylov 100 ~i ,.~/\l\\
.5
<l)
(1)
E 30 000 ' JI'!I :Im
.." 30,000
, if.,''~il,ii .
::> C' .,1' "
::> l\~''!\{\'~
Q
tJ 20,000
I '\

I 'f:/; .
);v'l:
1111
10,000 . j'!P~
~ ,~:,J~ I,
4~ll"

oa 0.1 0.2 0.3 0.4


Simulation time (ns)
(a)

Figure 4.
4.44 Magnetization reversal of a granular thin film element. (a) CPU time as a
function of the simulated time. (b) Transient magnetic states during switching.
switching,

At finite temperature the noise term has to be taken into account. As


shown by Garcia-Palacios and Lazaro (1998) the equation has to be
interpreted in the sense of Stratonovich, in order to obtain the correct thermal
equilibrium properties. The numerical integration of the stochastic differential
equation is performed using the method of Heun. For the pure deterministic
case the Heun method reduces to the standard second order Runge-Kutta
method (Kloeden and Platen, 1995). Numerical studies for simple spin
systems confirmed that the Heun scheme is numerically more stable and allows
larger time steps than the Euler or the Milshtein scheme (Scholz et al. ,
2001).
Micromagnetic Simulation of Dynamic and Thermal Effects 137

4.4 Numerical Examples

Section 4. 4. 1 presents the magnetization reversal dynamics of small


particles. Section 4.4.2 treats the hysteresis properties of granular thin film
elements. Section 4. 4. 3 compares different reversal modes of circular nano-
dots. The simulations in Sections 4.4. 1 to 4.4.3 are preformed neglecting
thermal fluctuations (T
CT =
0 K). The effects of thermal fluctuations on the
magnetization reversal of magnetic nano-wires is discussed in Section 4. 4. 4.

4. 4. 1
4.4.1 Small Particles
If a magnetic particle is sufficiently small it will reverse by uniform rotation.
Numerical experiments indicate that the dynamic properties obtained from the
simulation of a finite particle and a single magnetic moment are quite similar.
The switching dynamics were calculated for a columnar grain (Suess CSuess et al. ,
2001 a) as used in perpendicular recording (Richter,
2001a) CRichter, 1999) and compared with
the numerical results for a single spin. The geometry of such a grain can be
seen in Fig. 4. Sa.
5a. The basal plane of the irregular pentagon has a diameter of
12 nm. We varied the column length (the height of the grain), denoted below
by Ie. The material parameters are chosen for Co-Cr C (Po = 0.5
J.lo M s = O. S T, A = =
10- 11 Jim, K u = 3 X 105 J/m 3 ). ) . The easy axis is perpendicular to the basal

plane. The Gilbert damping constant is a =0.01. = O. 01. If the length of the particle is
smaller or equal to 20 nm the reversal process in uniform rotation. The
0
applied
external field is appl ied instantaneously at an angle of 1 l' off the easy axis.
Figure 4. Sb5b compares the calculated switching time as a function of the
applied field for the columnar particle and the single magnetic moment. The
solid line with circles in Fig. 4. Sb 5b shows the switching time for the columnar
grain with a column length of 20 nm and a damping constant a = = O. 01 as a
function of the field strength. We define switching time as the time until M z
crosses zero after the application of the external field. The switching times are
calculated for external fields in the range of about - 0.5 O. S x 2K u /CJ.lo
/( Po M s )
to-0.s
to-O. 5X2Kx 2K u /CJ.lo
/ ( Po M s )).' Since we neglect thermal activation in this
investigation, no switching is possible if I H ext I becomes too low.
Hex! It is
conspicuous that the switching time does not decrease with increasing external
field in the whole regime but shows a maximum slightly above the anisotropy
/(P o Ms ).
field, 2K u /CJ.l )' The switching time as a function of field strength for the
single magnetic moment (circles Ccircles in Fig. 4. 5)
S) strongly resembles that of the
columnar grain and helps to understand the dependence of the switching time
on the applied field.
138 T. Schrefl et al.

1 nm

E
c:
"
o
N

(a)
4.0
o T=OK
3.5 ~ Columnar grain
~
V>
- T=300K
5'"
S 3.0
<l)
Q)

E
f=
bJl 2.5
"c:
:.c:
:E
B
.~ 2.0
VJ
(/J

1.5

1.0
0 -1.0 -1.2 -1.4 -1.6 -1.8 -2.0
Hext[2KuI(poMs)]
Hext[2K,,I(J1oMs)]
(b)
Figure 4.5 Top: Small columnar Co-Cr particle. Bottom:
Bollom: switching time
lime as a function of the
field strength. Circles: single magnetic moment at zero temperature. Solid line: single
magnetic moment T=300 K. Solid line with circles: columnar grain at zero temperature.

Figure 4. 6 shows the energy as a function of the x and y component of


the magnetic moment just after the application of the external field of H Hex!
ext ==
- 0.9 X 2K uj (IJo
(/Jo M ss )) and Hex! = -- 1. 3 Xx 2K uj (IJo
Hext = (/Jo M ss )'
) ' respectively. Before

the application of the external field, the system is in equilibrium at Mx = My = 0


marked with a black dot in Fig 4. 6a and Fig 4. 6b. If a field is applied
instantaneously the energy landscape suddenly changes. The system is no
longer in equilibrium. In the case of Hex! =
Hex! = --1. 3X2K
1. 3 X 2K u ujj(/JoM
(IJo M ss )) a well defined
maximum is formed because the Zeeman energy dominates the total energy.
Due to the small damping constant in the LLG equation, the magnetization
moves along a path with almost constant energy around the maximum of the
energy surface. During this motion the angle between the magnetization and
the external field remains small. More precisely the angle. between M and
H ext (parallel to the z direction) is small during the initial motion of the
-- Hex!
magnetization. Thus the corresponding torque is small. The relaxation towards
the reversed state is slow. The switching time is long. If the external field is
comparable with the anisotropy field, the energy landscape is more complex.
The anisotropy energy and Zeeman energy contribute in the same order of
Micromagnetic Simulation of Dynamic and Thermal Effects 139

magnitude to the total energy. A path with nearly constant energy is no longer
a circle. The angle between M and - Hex! H ex, becomes higher (Fig. 4. 6a) which
leads to a high torque. The high torque leads to a fast relaxation towards the
reversed state. Fast switching modes are possible if the rise time of the
external field is shorter than the relaxation of the magnetization towards the
local minimum close to the initial state.

0.206
0.204
0.202
'"; 0.200
';
.;
oj
~0.198
;;; 0.198
0.196
0.194
0.192
0.19 I
0.4 0.2

(a)

0.40
0.39
0.38
,...,.
'";
=' 0.37
~
$
t<l
t<J 0.36
0.35
0.34
0.33 ;'
0.4 02
. 0
-0.2
MJ' -{).4
-0.4
(b)

Figure 4.6 Energy landscape as a function of M xx and My. The bold line shows the path of
the magnetic polarization after the application of an external field of (top) Hex! He,. =
-0.9 2K u/ (po M ss )) and (bottom) Hex!
X 2K,/(/Jo
- O. 9 X He,. = - 1. 3 X
X 2K,/(/Jo ). The black dots show the
2K u/ (Po M ss ).
initial state.

In addition, Fig. 4. 5 gives the switching time calculated for a single


magnetic moment at T = 300 K. The magnetic moment corresponds to a
magnetic particle with a volume of ((1212 nm)3
nm) 3 and a spontaneous polarization 110
IJo
Ms = O. 5 T. The results at finite temperature have been averaged over 100
M
simulations. The solid line in Fig. 4.5 shows that fast switching at low external
fields also occurs at 300 K. However, the finite temperature reduces the
140 T. Sehretl
Sehrefl et al.

switching time. In the investigated temperature range (0 - 400 K) the


linearly as a function of the temperature. The
switching time decreases almost Iinearly
fluctuation of the magnetization is responsible for the reduction of the switching
time at non-zero temperature. At non-zero temperature the number of
precessions until M =
Mz reaches zero is smaller than at T = O. This indicates that
the system at non-zero temperature is effectively stronger damped. The
thermal field causes fluctuations of the magnetic moment orientation. If the
applied field is almost parallel to the anisotropy axis, the thermal fluctuations
always increase the angle between the magnetic moment and the effective
field which causes a higher torque.

4. 4. 2
4.4.2 Thin Film Elements

Thin film magnetic elements are the basic structural units of magnetic memory
cells and magnetic sensors. A precise understanding of the switching process
and the possible tuning of the switching field and the switching time are helpful
for the future development of these devices. Magnetic nano-elements are found
to reverse by the formation of vortices, which in turn leads to the expansion of
the domain which has its magnetization parallel to the external field. Vortices
may nucleate from end domains which are formed in the remanent state to
minimize the magnetostatic energy (Kirk et al. , 1997).
1997>. Additional sources for
vortex nucleation are surface irregularities and grain boundaries. Numerical
micromagnetic simulations show that edge roughness and the polycrystalline
microstructure considerably lower the switching field of Co elements (Schrefl
et al. , 1999).
Figure 4.7
4. 7 gives the microstructure of the polycrystalline sample and
transient states during irreversible switching. The competitive effects of shape
and random crystalline anisotropy lead to a magnetization ripple structure.
Sharp edge irregularities help to create vortices, which will move through the
width of the element. This process starts at a reversed field of Hex! = - 95 kA/m
and leads to the reversal of half of the particle. In what follows, a second
vortex forms and the entire Co-element becomes reversed. In granular Co
elements with random magneto-crystalline anisotropy, vortices form
immediately after the appl application
ication of a reversed field. For zero
magnetocrystalline anisotropy a vortex breaks away from the edge only after a
waiting time of about 0.8
O. 8 ns. The coercive field of the flat element without
surface roughness and grain structure was found to be 140 kA/ m.
kA/m.

4.4.3 Circular Nanodots

The reversal process of circular nano-dots strongly depends on the diameter and the
thickness of the dots (Suess et al., 2001 b ). The larger the dots, the more
Micromagnetic Simulation of Dynamic and Thermal Effects 141

1$
-------,:-:-:------- ~

';;;'
5
0.6
'"
E
E=

1.13

Figure 4. 7 Microstructure of the polycrystalline sample and snapshots of the magnetization


distribution at different times after the application of an applied field of H ex, = - 95 kA/m.
He,'

nonuniform reversal modes are observed. In addition a small Gilbert damping


constant favors nonuniform reversal in comparison to a large damping constant.
Figure 4. 8 shows the finite element mesh at the surface of a circular nano-dot. The
= 0, J.lo
calculations were performed for NiFe (K u = = 1 T, A == 10
J..Io M s = 10-- 11 J/ m) .

/.~-;..
/.~~
~~~~
~~~~
~~~.
~~~
. '0_.;"
.

Figure 4.8 Top: Finite element model of one circular nano-dot. The triangles show the
surface mesh used for the boundary element method. Bottom: Three possible reversal
modes (rotation: diameter d = 55 nm, thickness t = 10 nm;
nm, nonuniform: d = 110 nm, t =
=
10 nm; nm. tt=
nm, vortex: d= 110 nm, = 15 nm).

Figure 4. 8 shows transient states during reversal for different sizes of the
magnetic nano-dot. A damping constant aex = 1 was used. The left dot has a
diameter of 55 nm and a thickness of 10 nm. For this small volume of the
particle the reversal process is homogenous rotation. The middle dot shows
the reversal process of a dot with d == 110 nm and a thickness of 10 nm. An
142 T. Schrefl et al.
a!.

s-state is formed which reduces magnetic surfaces charges and hence the
magnetostatic energy. If the thickness exceeds 15 nm, as in the right picture,
a vortex state has smaller energy than an s-state. Two reasons can be
mentioned why thicker samples favor the formation of a vortex state. First, in
the core of a vortex state the magnetization points perpendicular to the surface
and produces a demagnetizing field. With increasing thickness the
demagnetizing field decreases which reduces the magnetostatic energy.
Second, the surface charges at the cylindrical surface which lead to a high
magnetostatic energy become dominant with increasing thickness.
In addition, the Gilbert damping constant was found to influence the
reversal process. Decreasing the damping constant may change the reversal
mode from nonuniform rotation to vortex motion. Figure 4. 9 compares the time
evolution of the magnetic component parallel to the field for 0' = 1 and 0'ex ==
ex =
d = 220 nm and thickness t=
O. 01 (diameter d=220
0.01 t = 10 nm). After the application
appl ication of
a field H ext = -8 kA/m, 11 off the x-direction, the torque remains small. For
ex = 1 the nan-odot starts to switch only after a waiting time of about 3 ns. For
0'
0' =O. 01 the waiting time reduces to about O. 5 ns. Leineweber und KronmOlier
ex = Kronmuller
Kronmuller, 1999) observed that a certain waiting time is
(Leineweber and KronmOller,
required before switching is initiated in hard magnetic spheres. The insets
compare two transient states during reversal. Whereas for 0' ex = 1 the
magnetization reverses nonuniformly, two vortices are formed at the beginning
of the reversal process for 0' = O. 01 .
ex =
1.0
",- -a=O.OI
" ---a=l
\
\
\
\
0.5 \
\

\
\
\
\
~;. \
'i~<I~fT \
.4 .".4.4 <I' 4' \

o t~t~~t~tt:
.. \\.\lo).~. ~
\
\
\.)."\4"'}.+ ~ -.......... \
~..,..u.." \
\
\
o 4 6
Time (ns)

Figure 4.9 Time evolution of the magnetic polarization parallel to the external field
for different values of the damping constantC = 220, t == 10 nm) .
constant( d =

4. 4. 4
4.4.4 Magnetic Nanowires
The nucleation and reversal of reversed domains in magnetic nano-wires was
Micromagnetic Simulation of Dynamic and Thermal Effects 143

studied at non-zero temperatures. The energy barriers and the activation


volume were derived from the numerical results. The diameter of the Co wire
was 2 nm. The intrinsic magnetic properties of Co C = 1. 76 T, A == 1. 3 x
1-10 M s =
(lJo
1O- 11 J/m, K u = 6.88 x
=6. 5 3
X 10 J/m ) and a Gilbert damping constant a a == 1 were
assumed for the calculations. The magneto-crystalline anisotropy direction was
assumed to be parallel to the long axis of the wire.
Wires with a length smaller than 16 nm were found to reverse by uniform
rotation. Figure 4. 10 shows schematic diagrams of the energy barrier as a
function of the angle with respect to the easy axis. An external field lowers the
energy barrier. From the calculated relaxation time, T, the energy barrier can
be derived numerically fitting the numerical results to

T = , 1 (Es s
= f -0-1 exp k T ) (4.18)
C4.18)

'0
where f o is the attempt frequency. The energy barrier E s increases with
increasing volume of the wire as long as the reversal mode is uniform rotation.
The numerically derived energy barrier for uniform rotation agrees well with
the result obtained from the Stoner-Wohlfarth theory (Street
C Street and Crew,
1999) .

EsCH ext
Es(H KV(
ext )) = KV( 1 - ~e:t
%e;1 f, HKK= 2~sK
2':;sK (4.19)
C4.19)

where K is the effective anisotropy constant taking into account the


magnetocrystalline anisotropy and the shape anisotropy of the wire.

tAnis:tro
Anis:tro py axis

~Magnetization
VMagnetization

Angle
(a)

~
~I--------->r-----
w

(b)

Figure 4. 10 Energy of a small particle as a function of the angle between the easy axis
and the magnetization. (a)
Ca) zero external field and (b)
Cb) reversed applied field.
144 T. Sehretl et a!.
al.

Wires with a length of 32 nm reverse by the nucleation and expansion of


reversed domains. Now the energy barrier is independent of the wire length,
since the nucleation process starts at the wire and ends where a strong
demagnetizing initiates magnetization reversal. Once the reversed domain has
formed it expands over the entire wire. Figure 4. 11 illustrates this process.
An effective activation volume, v, can be derived from the energy barrier
under the assumption that the activation energy corresponds to the energy of
the nucleus of reverse magnetization (Street et al. , 1999)

E B ( H ext) =- VIJ 0 M s H ext (4.20)


The activation volume v can be derived from the slope of the E B (H ext) curve
1 () E
dEB
v
V = - - - - -B. (4.21)
() H ext
IJo M s d
JJo

.. 64 nm

Figure 4. 11
-
Nucleation and expansion of reversed domains in Co nanowires with an
aspect ratio of 1 : 16.

5 \
\
\
\
4 \
Wire(16:1 \
\
\
\
\
\
\
\
\
\
\
\
\
"-\ ....
o'--__
o'---__ ~~ __ ~ ~c.,__ _~
~c..,_ "
600 800 1000 1200 1400
External field (10 3 Aim)

Figure 4. 12 Energy barrier as a function of the applied field. The open symbols give the
numerical values. The dotted line is a linear fit of the numerical values. The dashed line
gives the analytic result according to (Braun 1994).
Micromagnetic Simulation of Dynamic and Thermal Effects 145

Figure 4. 12 gives the calculated energy barrier for a 1 : 16 aspect ratio nano-
wire as a function of the external field. The dashed line corresponds to the
analytical result obtained by Braun (1994). The numerical values for the
energy barrier are about a factor of 2 - 3 smaller than those derived
analytically. This may be attributed to inhomogeneous magnetic states which
are neglected in the analytical model. The numerically-obtained energy barrier
depends Iinearly
linearly on the applied
appl ied field, which indicates that the reversal takes
place by the formation of a nucleus of reverse magnetization at one end of the
wire. The activation volume was derived to be v = 2. 13 nm3 which
approximately corresponds to the cube of the wire diameter. Li and co-
workers (Li et aI., 1997) obtained a similar result from magnetic
measurements on ex-Fe nanowires.

References
Braun, H.-B. J. Appl. Phys. 76: 6310 (1994)
Brown, W. F. , Jr. Micromagnetics New York, Wiley (1963)
F. , Jr. IEEE Trans. Magn. 15: 1196 (1979)
Brown, W. F.,
Dahlberg, E. D.and J. G. Zhu. Physics Today 48: 34. (1995)
Fredkin, D. R. and T. R. Koehler. IEEE Trans. Magn. 26: 415 (1990)
Garcia-Palacios, J. L. and F. J. Lazaro. Phys. Rev. B 58: 14,937 (1998)
Gilbert, T. L. Phys. Rev. 100: 1243 (1955)
KronmOlier. IEEE Trans. Magn. 34: 3992 (1998)
Hertel, R. and H. Kronmuller.
Hindmarsh, A. C. and L. R. Petzold. Computers in Physics 9: 148 (1995)
Kikuchi, R. J. Appl. Phys. 27: 1352 (1956)
Kirk, K. J. , J. N. Chapman and C. D. W. Wilkinson. Appl. Phys. Lett 71:
539 (1997)
Kloeden, P. E. and E. Platen. Numerical Solution of Stochastic Differential
Springer Berlin, Heidelberg (1995)
Equations SpringerBerlin,
KronmOlier. J. Magn. Magn. Mater. 192: 575 (1999)
Leineweber, T. and H. Kronmuller.
Li, F. L., R. M. Metzger and W. D. Doyle., IEEE Trans. Magn. 33: 4423
( 1997)
Mansuripur, M.J. Appl. Phys.63: 5809 (1988)
Prinz, G. A. J. Magn. Magn. Mater. 200: 57(1999)
Ramstock, K. , T. Leibl, A. Hubert. J. Magn. Magn. Mater. 135: 97 (1994)
Rave, W. , K. Ramstock and A. Hubert. J. Magn. Magn. Mater. 183: 329
(1998)
Richter, H. J. J. Phys. D: Appl. Phys. 32: 147 (1999)
Schrefl, T. J. Magn. Magn. Mater. 207: 45 (1999)
Kirk and J. N. Chapman. J. Appl. Phys. 85: 6169
Schrefl, T., J. Fidler, K. KirkandJ.
(1999)
Scholz, W. , T. Schrefl and J. Fidler. J. Magn. Magn. Mater. 196 - 197 :
146 T. Schrefl et al.
at.

933 (1999)
Scholz, W., D. Suess, T. Schrefl and J. Fidler. Computational Materials
Science 18: 1 (2000)
Scholz, W., T. Schrefl and J. Fidler. J. Magn. Magn. Mater. 233: 296
(2001)
Street, R. and D. C. Crew. IEEE Trans. Magn. 35: 4407 (1999)
Suess, D. , T. Schrefl and J. Fidler. IEEE Trans. Magn. 37: 1664 (2001a)
Suess, D.,
D. , T. Schrefl, J. Fidler and V. Tsiantos. IEEE Trans. Magn. 37:
1690 (2001b)
Terris, B. D. , L. Folks, J. E. E. Baglin, A. J. Kellock, H. Rothuizen, P.
Vettinger. App. Phys. Lett. 75: 403 (1999)
Yang, B. and D. R. Fredkin. IEEE Trans. Magn. 34: 3842 (1998)
Zhu, J. -G. and H. N. Bertram. J. Appl. Phys. 66: 1291 (1989)

Work supported by the Austrian Science Fund (Y132 PHY, P13260-TEC).


5 Magnetic Relaxation and Quantum Tunneling of
Magnetization

X.X.Zhang

s. 1
5. Introduction

Magnetic relaxation effect is becoming more and more evident and increasingly
crucial in nanostructured magnetic systems as the size of the particles (or
clusters) decreases. Therefore, to understand the physics of the magnetic
relaxation is fundamentally important for both basic science and industrial
applications (Dormann et ai.,aI., 1997), which is actually one of the most
important issues in the high-density magnetic recording media. It has been well
known that magnetic relaxation effect is due to the magnetic moment flipping
caused by thermal energy. The flipping frequency of the magnetic moment in a
magnetic small particle, r, I, can be obtained from the Neel model (Neel,
1949a, 1949b):

rI = vexP(;B ~), (5.1)

where v is the attempt frequency of the order of 10 10 - 10 13 Hz (Johansson et


ai.,
aI., 1993; Linderoth et ai.,
aI., 1993; Prene et ai.,aI., 1993; Dickson et ai.,
aI.,
1993), U = = KV is the anisotropy energy barrier (K is anisotropy constant,
and V is a volume), k B is Boltzmann constant, and T is the temperature in
Kelvin. Actually the exponential relaxation law had been established
establ ished before
the Neel model (Becker and Doring, 1939). Since the anisotropy constant is a
material property (also related to the shape and the size of the particles), it
cannot be infinitely large. As recording density increases (V deceases), rI
will increase exponentially, and eventually the recorded information will
become unstable. To suppress the magnetic relaxation effect or to keep the
information, the magnetic recording media should be kept at a low
temperature. Suppose that the recording media is kept at a very low
temperature, for example, liquid He temperature; does the magnetic
relaxation effect completely vanish? It has been demonstrated that the
relaxation of magnetization is still happening with a constant rate even at such
a low temperature (and below), at which the relaxation effect due to the
148 X.X.Zhang

thermal agitation is completely negligible. The temperature independent


magnetic relaxation below a certain temperature is ascribed to quantum
tunneling of magnetization, which has been one of the most interesting topics in
the condensed matter physics since the beginning of the 1990 (Chudnovsky
CChudnovskyand
and
Gunther, 1988; Awschalom et ai.,
al. , 1990, 1992; Stamp et al.,
al. , 1992; Zhang et
al. , 1992; Tejada et ai.,
al., aI., 1993a, 1993b; Barbara et ai.,
aI., 1993; Kodama et
al. , 1994; 0' Shea and Perera, 1994; Ibrahim et ai.,
aI., 1995; Sappey et al. ,
1997>.
1997). It is, therefore, of great importance for both the fundamental research
and information storage technology to study the relaxation effect caused either
by the thermal agitation or quantum tunneling effect in the nanostructured
materials. In this chapter, we will discuss the physics and several
experimental evidences related to the thermal and quantum relaxation of
magnetization in nanostructured magnetic systems.

5.2 Magnetic Relaxation and Related Phenomena in


Monosized, Non-Interacting Particle Systems

5.2.
5. 2. 1 Introduction

For a magnetic single domain particle with a volume V and a uniaxial


anisotropy (the anisotropy constant is denoted by K u (J/m 3 , )), the
magnetization M o = M s V points either "up" or "down" in a zero field, if the
easy axis is along the z-axis. In the picture of energy spectrum, the "up" and
"down" directions of the magnetic moments are the double degenerate states,
as shown in Fig. 5. 1.

Figure 5. 1 Double degenerate states for magnetic moments


moments""up" and "down.
up" and" down. "

The rate for M o changes from "up" to "down" (or "down" to "up") and is
(5. 1 and energy barrier
governed by the exponential law (Eq. (5.1))

U = K u V, (5.2)

if the flipping of M o is caused only by the thermal activation. It is evident that


Magnetic Relaxation and Quantum Tunneling of Magnetization 149

the flipping rate (/) decreases exponentially with decreasing temperature


(T). At very low temperatures, where U/k Bs T30, M o will stay "up" or
"down" for quite a long time. The relaxation time (T), the average time for
M o changing the direction, can be estimated as T= 1//= 1O-10exp(30)s~
10 3 s. At a high temperature, for example, UU// k Bs T ~ 10, the relaxation time
6
becomes as small as '" _10-
10- s. In this case, one will see that the magnetic
moment flips very fast. This behavior of the moment is actually very similar to
that of the moment in a paramagnetic free ion. Since the magnetic moment in a
single domain particle (for example, 5 x 1044 JJs
JJ.B for a Co particle of 4 nm) is
much larger than that of a magnetic ion (the maximum is 10.64 JJs JJ.B for Dy
Dy3+
3+ ),
),

this paramagnetic-like behavior is called "superparamagnetism."


"superparamagnetism. "

5.2.2
5. 2. 2 Blocking Temperature

As we discussed above, the relaxation time for a magnetic moment changes


drastically with varying temperature. What will one observe if a technique is
used to continually monitor the behavior of the moment? Suppose we use a
SQUID magnetometer to measure the magnetic moment and the time for the
SQUID magnetometer to take one measurement is t mm.. If the relaxation time of
Mo is much longer than t m , one will find Mo being fixed "up" (or "down").
When temperature is high enough, M o changes its direction many times in the
period t mm', showing a superparamagnetic behavior. At a certain temperature
B , the relaxation time of the M o is equal to (or comparable with) t m
T s' m', one
will "see" M o"up" or "down" with almost the same probability, if one takes
many measurements continuously at this temperature. Just below T s' B' the

probability to "see" M o in one fixed direction, for example "up" (or "down"),
is much larger than that to "see" it "down" (or "up"). If the temperature T is
just above T sB ', one will find that M o is always flipping. The temperature T Bs is,
therefore, called blocking temperature. We can estimate easily the value of
T sB by using the characteristic measuring time t m' the value of U and the
relation
(5.3)

We find
U = In ( vt m) k s T s . (5.4)

For a typical value of v = 10 10 S -1


-1 and the measuring time t m = 60 s (for SQUID

magnetometer), Eq. (5. 4) becomes

=
U = 25k Bs T Bs (5.5)

It is evident that the blocking temperature is proportional to the size of an


energy barrier. One should also note from Eq. (5.4)(5. 4) that the blocking
150 X.X.Zhang
X. X. Zhang

temperature could be very different for different techniques with very different
measuring times t mm . The values of t mm may change from 5 x 10- 9 s for
M6ssbauser spectroscopy (M0rup and Tronc, 1994) to 10 2 s for some d. c.
Mossbauser
magnetic measurements (for example, SQUID magnetometer). Especially for
ac susceptibility measurement, the frequency of the a. c. field can be changed
from 1O- 3 3
s- 1 to 10 44s- 1 , i.e., 1O-
1O-4s~tm~103s.
4
s":;;;t m ":;;;10 3 s. We now use an example to
demonstrate how the values of t maffect m affect T B' For a 4 nm Co particle, the values

of T BB are calculated by using magnetic anisotropy constant K uu =5.3 = 5. 3 xX 10 5 J/m 33


(Craik, 1995), v = 10 1s- 0s- 1 and Eq. (5. 4). As shown in Fig.5. 2, the blocking
temperature changes from ....... 330 K for the Mossbauer
-330 M6ssbauer spectrum to <50 K for
ac susceptibility measurement with a very low frequency. Therefore, when talking
about the blocking temperature, one should also tell which technique is used.
350
-Mossbauer
300

250

~ 200
~
~
150

100 SQUID

I
50
ac susceptibility
oO~~~"""""''''''''''-=---'---e-~-,--,-:,.........,.-'-~e...,.-....".........."...........,
c,......~'""'::'-......."..........,...........,............".........,:"""-'.,......".......'-:'-...."........."............,........
10-910-810-710-610-510-410-310-210-110 101102103104105 10 1 10 2 10 3 10 4 10 5
.......

Measuring time tm,(s)

Figure 5. 2 Blocking temperature as a function of measuring time t1m (solid line).


The time ranges for different techniques are indicated by the symbols.

5.2.3 ac Magnetic Susceptibility Measurement

From Eq. (5. 4) and Fig.


Fig.5. 2, it is known that the value of T B is determined by
5.2,
the measuring time t mm and the energy barrier U. This property has usually
been used to obtain some parameters of the system, e. g. attempt frequency
and energy barrier, etc., by varying t m(M0rup
m (M0rup and Tronc,
Trone, 1994; Zhang,
2000) . For this purpose, temperature-dependent ac susceptibility data
2000).
measured with different frequencies are commonly used. The ac susceptibility
measurements are usually performed with zero dc field. The amplitude of the
ac field is usually very small (less than a few Oe), or small enough not to
affect the intrinsic energy barrier in the system. When an ac field with a
Magnetic Relaxation and Quantum Tunneling of Magnetization 151

frequency f is applied to a magnetic system, the magnetic susceptibility is


given as (Mydoch, 1993; Fannin et al.
al.,, 1993)
X ( w) = X' ( w) - i X" (w) . (5.6)
(5. 6)

The in-phase (real part) susceptibility X' and the out-phase (imaginary part)
susceptibility X" are related by the following equations,

'_ + XT - Xs (5.7a)
X - Xs 1+ (WT)2

=
X" = WT (
WT( 1~T ~W~;
1~ ~W~)2)
2 )
(5.7b)

where T is the relaxation time given by T = 1I r = v --II exp( V I k B T), X


XTT is the
isothermal susceptibility in the limit w-O (or dc susceptibility), and Xs XS is the
w-oo (or very high frequency). From Eq.
adiabatic susceptibility in the limit W-OO
(5. 7b), it is evident that there are two possibilities for X" = 0, i. e. , when
WT = 0 and WT = 00. When WT = 0, one measures the equilibrium isothermal
susceptibility XT i. e., X = XT; whereas at WT = 00, one measures the
nonequilibrium adiabatic susceptibility, X = X Xss (Mydoch, 1993). Another
important fact that should be noted from Eq. (5. 7) is that when WT = 1, X"
shows a maximum. For a given temperature, this maximum provides a method
for determining the relaxation time T (Mydoch, 1993).
For a single domain magnetic particle, when the temperature increases
from a very low temperature (the blocked state) to a high temperature
(superparamagnetic state), the value of the relaxation time T changes from a
very large number to a very small one. Therefore, for a given ac field
(w = 2nf in Eq. (5.7)),
frequency f (w=2nf (5.7, WT will decrease from 00 (at very low
T) to 0 (at very high T). Consequently both X' (T) and X" X"((T)
T) will show a
maximum. The peak appears in X"( T) at WT = = 1, i. e., w[ V-I exp( VI k BB T p ) ]
= 1, where T p demotes the peak temperature in x"
= X" (
( T). By changing the
frequency f, the peak temperature will shift accordingly. This relation can be,
therefore, used to extract the energy barrier height and the attempt frequency,
by plotting the data of frequency and the peak temperatures using the following
relation:
V
In2nf = In( v) - k T ' (5.8)
B B

T data obtained from


Figure 5. 3 shows the ac susceptibility (X' ( T) and X" ( T))
a magnetic fluid composed of mono-sized Co nanoparticles of 4 nm in diameter
(Zhang et al.
a!. , 2002). Each Co particle is coated with an oleic acid layer of 2
nm in thickness that serves to stabilize the particle and prevent oxidation
(Woods et a!.
al. , 2001). To avoid the dipole-dipole interaction and to have a
signal large enough for the ac susceptibility
susceptibil ity measurement, we made the
152 X. X. Zhang

magnetic fluid with a volume concentration of~O. 3 vol. %. It is clearly seen


that with increasing the frequency f, the peaks in both X' (T) and X" (T) shift
to higher temperatures. It is found that the peak temperatures T ~ s in X" ( T)
v = 1. 0 X 10 14 S-1
can be well fit to Eq. (5. 8), with v= s- 1 and U/k = 733. 4 K5.
U/ k ss =733.4 K 5. 69 K.
The extracted value of energy barrier corresponds to an anisotropy constant K
== 3. 02 X 10 5 J/m3 for a 4 nm particle,
particle. which is consistent with the K K== 1. 87 xX
10 5 J/m3 and K K = 3.08= X 10 5 J/m3 for 5 nm and 3 nm Co (mt-fcc) particles
3. 08 X
(Woods et al. , 2001). These values of the anisotropy constant were obtained
by measuring the temperature-induced spontaneous magnetic noise of
nanoparticle films that were composed of self-assembled lattices of uniformed
superparamagnetic nanoparticles. The magnetic-noise study is actually based
on the same physics as that for the ac susceptibility measurement. Some other
values of K for fcc-Co were reported, 2.7 x 105 J/m3 (Sucksmith and Thompson. Thompson,
1. 2 x 10 5 J/m3 (Chuang et al.,
1954), 1.2X a!., 1994) and 5X 5 x 10 5 J/
J/mm3 to 3x
3 x 106 J/
J/mm3
aI., 1995). Thus, our K value is in general agreement with the
(Chen et al.,
values reported.
1.2

..,
~
14
12 ~
1.0 n. l \'\'*
E'~JPi
0'" ~\ ~
Q) Q)

~ 10
-=~E
0
0.8
-=
~ 8
E /;~ "J,'.\
if" ,

6'"
b
xX
~
66
b0
on'"
I

X
~
;:-'
Q)

0.4
0.6
I \\\~
l--O.1 Hz
~\\
.:: \ \

"':;;
<-< 4 ~~ '0'0"".
/-0-0.3 Hz
2 0.2 - A - 3.0 Hz ~~.
-*- 33 Hz ,,~
0o 20 40 60 80 100 o0 20 40 60
T(K) T(K)
Figure 5.3 Temperature dependent magnetic susceptibility obtained on Co nanoparticles.

One should note that the fitted attempt frequency.


frequency, '" -- 1 x 10 14 SS -1.
-1, is

significantly larger than the values reported (Johansson et al., 1993;


Linderoth et al., 1993; Prene et a!.,
alo' 1993; Dickson et al., aI., 1993).
1993) This 0

attempt frequency corresponds to a relaxation time of 10 -14 s, which is shorter


than the shortest time scale in magnetism,
magnetism. T 5 '"
5 - - 10 -13 S for the spin fl ip time of

a single atom. Since no time scale should be shorter than T s' s ' the Arrhenius
law may not be the proper law to describe the physics of the relaxation peak
a!., 1996; Djurberg et al.,
(Zhang et al., aI., 1997; Jonsson et al., 1995). It is
generally believed that two factors in the particles system will lead to non-
Arrhenius law behavior. The first factor is the size distribution. It may not be
important in our sample, because it is composed of mono-dispersed particles.
The second one is the dipole-dipole interactions between the particles, which
can certainly change the magnetic behavior of the particles system, and even
Magnetic Relaxation and Quantum Tunneling of Magnetization 153

a spin-glass behavior can be observed from some systems with strong


interaction (Luo et al. , 1991; M0rup and Trone,Tronc, 1994; Jonsson et aI., al. , 1995;
al. , 1997; Garcia-Otero et al. , 2000).
Djurberg et al.,
The dipole-dipole interaction can be estimated by using E s = (Pol
d- d/ k B
Ed-dlk (JJo /
4'TTk s)
4rrk M;
B) M; VE (Djurberg et al., 1997), with the saturation of magnetization
=
Ms = 1 435 emu/cm 33 (Cullity, 1972), V = 3.34= =
3. 34 X 10- 26 m3 and E = 0.3 %. The
calculated value for the interaction is (E d-d/ =)
B = ) 1. 5 K. The interaction
d- d I k s

effect can also be estimated by fitting the equilibrium susceptibility data to the
Curie-Weiss law X 0 = c C I/ ( T - e). The fitting gives e= - 5. 2 K, indicating an
antiferromagnetic interaction. For the interacting systems, the Vogel-Fulcher
law,

T = TOexP[k s ( T ~ To) ] (5.9)

is frequently used to describe the physics of the relaxation peak, where To


indicates the strength of the interaction. By fitting the values of T p to the
Vogel-Fulcher law and taking To = = 5.2
5. 2 K, we obtained UI = 447 K 25 K,
U/ k Bs =
1. 9 X 10- 11 s. This value of To has now physical meaning and is in
To = 1.9
agreement with those used in other particle systems (Djurberg et al. , 1997).
The value of energy barrier U I k sB = 447 K corresponds to an anisotropy
U/k
constant of 1. 85 x 10 5 J/m3 , which is very close to that obtained for 5 nm Co
particles by using the magnetic-noise experiment (Woods et al. , 2001). The
fact that the interaction between the particles obtained from the ac
susceptibility (5. 2 K) is larger than the calculated value (1. 5 K) may suggest
that the diluted sample is not completely homogeneous and that there exists
small agglomerates of particles in this sample.
Another piece of information that can be obtained from the ac
susceptibility data is the energy barrier distribution. It is well known that, for a
single barrier system, if the frequency range of the ac field is broad enough,
the Argand diagrams will be a half-circle (Mydoch, 1993; Zhang, 2000). If
the barrier is independent of temperature, the Argand diagrams measured at
different temperatures will collapse into one master curve. That means that if
there is an energy distribution caused either by the size distribution or by the
dipole-dipole interaction, the Argand diagrams will deviate significantly from
the half-circle. Therefore, the deviations from the half-circle may serve as an
estimation for the energy distribution (Mydoch, 1993). Now let us look at the
Argand diagrams obtained from the magnetic fluid used for the above ac
susceptibility study. Towards this aim, we have performed the ac
susceptibility as a function of frequency at different temperatures just below the
block ing temperature.
blocking
Choosing these temperatures is due to the limitation of our frequency
3
(10 - 22 -10
range (10- - 10 Hz). If the temperature is too low or too high, the imaginary
154 X.X.Zhang
X. X. Zhang

part will be very small. The Argand diagrams obtained at T = 26, 27, 28 K
are shown in Fig. 5.4.
5. 4. It is evident that all the data collapse to a master curve
and the Argand diagrams can be fitted to a circle, indicating a very narrow
distribution of relaxation time (or energy barrier). The non-full-half-circle in
Fig. 5.4
5. 4 could be due to a non-single barrier property of the system caused
either by a narrow distribution in volume (Woods et al., 2001) or by the
interaction discussed above. It could be also due to our frequency range being
too narrow.
1.4

vV' 1.2
~
E 26K
10'"
'f'v o 27K
27 K
1.0 .. 28K
X
'-'
~

.~
- Full Half-Circle
'~

0.8

0.4 0.6 0.8 1.0 1.2 1.4 1.8


X'(X I0-44 emu/Oe)
X'(XIO- emu /O e)

Figure 5.4 The Argand diagrams for Co particles. Symbols are the data obtained at
different temperatures; the line is the fitted full half-circle.

So far, a perfect full-half-circle Argand diagram has not been observed in


nanoparticle systems, due to the energy distribution caused either by size
distribution or by dipolar-dipolar interactions. Although the particles used in
this study are already very uniform and the magnetic fluid is very dilute, there
is still a relaxation time distribution. The only exceptions are the magnetic
molecule single crystals, e. g., Mn12 and Fes aI., 1996;
Fea (Friedman et al.,
Hernandez et al., 1997; Zhang et al., 1999). In these magnetic molecular
single crystals, all magnetic clusters are identical and have a spin (S) of 10.
In addition, these clusters have a very strong magnetic anisotropy (U / k B ~
60 K and 30 K for Mn12and
Mn12 and Fes
Fea respectively) and their easy axes are along the
same direction. As such, these molecular single crystals represent an ideal
model system with a single barrier, which can be used' to examine
experimentally some fundamental physics and to look for new physics (Zhang
et al. , 2001). Since the anisotropy energy is high (U / k B ~60 K for Mn12) and
the spin is small (S = 10), the energy difference between the spin levels is
appreciable, which leads to the discovery of the resonant quantum tunneling
tunnel ing
(Friedman et al.al.,, 1996; Hernandez et al.,
al. , 1996; Thomas et aI.,
al. , 1996). The
fundamental characteristics of Mn12 molecules have been reviewed
(Chudnovsky and Tejada, 1998). Since all the spin levels in the Mn12 are in
Magnetic Relaxation and Quantum Tunneling of Magnetization 155

resonance at zero magnetic field, the energy barrier extracted from the ac
susceptibility data (with zero dc field) may not be due to only one spin-level
(Hernandez et ai.,
aI., 1997). It is also known that the dominating spin level in
tunnel ing process depends strongly on the temperature (Hernandez et al. ,
tunneling
1997), or in other words, the energy barrier will change with varying the
temperature. To have a temperature-independent, single barrier, a proper
field (3.3 kOe) is applied to mismatch the spin level in both sides of the
energy barrier, which prevents the resonant tunneling effect (Zhang, 2000).
All the spins flip through the thermal activation, which leads to a full-circle
Argand diagram, Fig. 5. 5.55 (Zhang, 2000). It is actually a first observation of a
perfect half-circle Argand diagrams in magnetic system.

3.0 H=3.3 kOe


o 0

.,
2.5

~
2.0
a
'i'"
I' 1.5
S
.
~
2S
~
~
1.0
o

5.6K
6.8K
- - Full half-circle
0.5

0.0
I 2 3 456 7 9
X'(XIO-5emu/Oe)
X'(XIO-semu/Oe)

Figure 5.5 Mn12 clusters obtained at H = 3.3 kOe. Symbols are


The Argand diagram for Mn'2
the data obtained at different temperatures; the line is the fitted full half-circle.

5.2.4 Zero-Field-Cooled and Field Cooled Magnetization Curves

The temperature-dependent magnetization data measured in zero-field-cooled


(ZFC) and field-cooled (FC) procedures are usually used to obtain the
information of the energy barriers. The ZFC-FC magnetization measurement is
carried out as follows. For the ZFC magnetization curve, the sample is first
cooled in a zero field from a high temperature well above T B' where particles
are in a superparamagnetic state, down to a low temperature well below T B .
Then a magnetic field is applied and the magnetization as a function of
temperature is measured in the warming process to a temperature well above
the blocking temperature. The FC curve is obtained by measuring the
magnetization when cooling the sample to the low temperature in the same
field. If the magnetization cannot be measured in the cooling process due to
156 X.X.Zhang

technical difficulty, the FC curve can be obtained in an alternative way. That


is, cool down the sample to the low temperature in the same field and then
measure the magnetization with increasing the temperature. In the ZFC and FC
measurements the field must be weak enough in comparison with the
anisotropy field to guarantee that the ZFC-FC curve reflects the intrinsic
energy barrier distribution.
Let us now discuss why useful information can be obtained from the ZFC-
FC magnetization data. It should be noted that the net magnetization is zero. zero,
i. e. the sample is at the equilibrium state.
state, before a magnetic field is applied
to the sample in the ZFC measurement. After a low field is applied, appl ied, the
equilibrium is broken due to the Zeeman energy. If the particles are
isotropic. i. e., there is no energy barrier involved in the
magnetically isotropic,
magnetic moment flipping processes, the curve is a measure of the
thermodynamic equilibrium states being well described by the Curie law.
When energy barriers exist in the system. system, the sample cannot reach the
corresponding thermodynamic equilibrium state at such a low temperature and
a low field, and the system will try to approach the corresponding equilibrium
state slowly through relaxation processes. With increasing the temperature,
the relaxation rate increases exponentially and the magnetic moments of more
and more particles will turn to the field direction, i. e.,
e. , the magnetization of
samples increases with temperature. At a certain temperature, the
magnetization increases to the value of the equilibrium state. If the
temperature increases further, it then decreases following the Curie law, and
consequently a maximum appears at this temperature. This peak temperature
is normally called the blocking temperature T s' B' which can be estimated by

equalizing the measuring time t m m to the relaxation time of the magnetic


moment, as described above. The FC curve should be well described by the
Curie law particularly for T> T B' s ' in the thermodynamic equilibrium states. If
the cool ing process is too fast, the system appears to be "quenched."
"quenched," the
<
measured FC data for T < T Bs no longer reflect the thermodynamic equilibrium
states. This situation is normally observed in the FC measurement by cooling
the sample directly to the low temperature and then measuring FC in the
warming process. From the ZFC curve we can obtain the energy barrier height
by using the relation given by Eq. (5. 4) and the energy barrier distribution
from the width of the peak.
In Fig. 5.6, the ZFC-FC magnetization data obtained from the same
magnetic fluid used for ac susceptibility measurements (shown in Fig. 5.3) are
plotted as a function of temperature. The energy barrier height was calculated
81 0 K, by using the blocking temperature T Bs = 30 K, v = 1/ To ( To
as U / k Bs = 810
=1.9X10- 11 s), tm=10s, and U=ln (vt m )k Bs T Bs (Eq.(5.4.
(Eq.(5.4)). This value is
much larger than the value obtained from the ac susceptibility data. This
deviation may be caused by the dipole-dipole interactions and some
Magnetic Relaxation and Quantum Tunneling of Magnetization 157

aggregations in the sample. In fact, two techniques give the information of the
magnetic response of the system in a very different time scale. A single ac
measurement gives a response of the particles in the system with the flipping
time being just comparable to 1/ w. The particles with flipping time is much
larger or smaller than 1/ w, and give no contribution to the X" (the imaginary
susceptibility). But, the ZFC magnetization is a result of the accumulated
flipping processes. When there is an energy distribution or a dipole-dipole
interaction, both can only give an average energy barrier height rather than
the exact value, which may lead to a larger value extracted from ZFC data
than that from ac susceptibility data.

0.0020 Fe

0.0016
-e-
- . - Magnetic fluid
- 0 - Dried sample
]' 0.0012
10.0012
~
~ 0.0008 J'
pO
ZFC ?
0.0004
~
I TB=40K

0.0000 TB=30 K

oo 20 40 60 80 100 120 140 160


T(K)
Figure 5.6 The ZFC-FC magnetization curves obtained on systems composed of mono-
sized, 4 nm Co nanoparticles. Solid symbols: diluted magnetic fluid with concentration <
0.3 vol%; open symbols: dried magnetic fluid.

The effects of magnetic dipolar interaction have been intensively


investigated experimentally (Dormann et aI., 1988; Luo et aI., 1991; M0rup
and Trone, 1994; Jonsson et al. , 1995, 1998; Djurberg et al. , 1997; Mamiya
et aI., 1998;) and theoretically (Shtrikman and Wohlfarth, 1981; Chantrell
and Wohlfarth, 1983) or numerically (Garda-Otero et al. , 2000), for more
than ten years. Most of the experimental studies have been carried out by
using concentrated (dense) magnetic fluids. The techniques used in the
mentioned experiments include Mossbauser spectroscopy, ac susceptibil ity
susceptibility
measurement, and dc de magnetic measurements (ZFC-FC and magnetic
relaxation), etc. Different experimental results were obtained, e. g., by
increasing the particle concentration (or increasing the interaction) the
effective energy barrier associated with the magnetic moment flipping
increases in one experiment (Dormann et aI., 1988), whereas the barrier
decreases in the other system (M0rup and Trone, Tronc, 1994). Another very
interesting question is Whether or not the interaction can lead to a spin-glass
158 X.X.Zhang
X. X. Zhang

(Luo et al. , 1991; Jonsson et al. , 1995, 1998; Mamiya et al. , 1998; Garcia-
Otero et al. , 2000). Experimentally, some characteristics of the spin-glass
have been observed in the concentrated frozen magnetic fluids (Luo et al. ,
1991; Jonsson et ai.,
al., 1995, 1998; Djurberg et ai., al., 1997; Mamiya et al. ,
1998). The analysis on the field dependent ac susceptibility data by using the
static scaling showed a divergent behavior of nonlinear susceptibility, which
strongly supports the law-temperature spin-glass-like phase in the interacting
nano-particle systems (Jonsson et al. , 1998). However, the recent numerical
simulation results indicate that no spin-glass phase can be formed by the
dipole-dipole interaction in particle assembly, although the energy barrier
increases (or the blocking temperature increases) with increasing the dipole-
dipole interaction by reducing the distance between the particles (Garcia-
Otero et al. , 2000).
The dipole-dipole interaction effect was also studied by using the 4 nm Co
particles. To reduce the distance between the particles, the magnetic fluid
sample was dried slowly in Argon gas. The Co nanoparticles should self-
assemble into 3D close-packed arrays with a lattice parameter of about d +
4 nm where d is the diameter of the particles (Woods et al. , 2001). In this
case, d = = 4 nm, which corresponds to a volume concentration of about 9
vol. %. The dipole-dipole interaction between the particles was calculated to
be 46 K by using Ed-dlk = (po/4rrk
E d - d / k Bs = = 9.2 %. Let us now have
M; VE and E =9.2
(/..Io/41Tk Bs ))M;
a look at how the increased interactions affect the magnetic behavior of the
sample. In Fig. 5. 6, the ZFC -FC magnetization curve for the dried sample is
plotted with that for the liquid sample. It is clearly seen that the peak shifts to
a much higher temperature (from 30 to 40 K), indicating that the dipole-dipole
interaction will increase the energy barrier. This feature is consistent with the
other experimental studies (Jonsson et al., 1995, 1998; Djurberg et al. ,
1997;) and numerical study (Garcia-Otero et al. , 2000).

5.2.5 Magnetic Hysteresis Loops


As we discussed above, at temperatures T T s' B' it is very difficult for a

particle system to reach its thermodynamic equilibrium state in a field much


less than the anisotropy field H K due to the slow relaxation process. With
increasing the applied field, magnetic moments of more and more particles
turn to the field direction through the relaxation process. At a certain field, the
magnetization of the system gets saturated. The saturation field is always
smaller than the anisotropy field at a finite temperature, which decreases with
increasing temperature. This is due to the fact that, at higher temperatures,
the relaxation rate becomes exponentially large and the system can reach the
thermodynamic equilibrium easily. Actually, when the temperature is higher
Magnetic Relaxation and Quantum Tunneling of Magnetization 159

than the blocking temperature, the magnetization curve can be well described
by the Langevin function. At T < T B' as the applied field decreases from the
field higher than the saturation-field, the demagnetization will follow the
magnetizing curve down to the saturation field H s . By further decreasing the
field, the demagnetization curve will not follow the magnetizing curve, and
instead be higher than it, provided that the sweep rate of the field is not
infinitely slow. This can be understood as the following. Since the slow
relaxation processes dominate the magnetic moment rotation, the system
cannot respond immediately to the field change to reach its thermodynamic
equilibrium state described by the Langevin function, which leads to a
magnetic hysteretic behavior. Thus, when the field is decreased to zero, the
magnetization will decrease to a finite value, the remnant magnetization, Mr.
Increasing the field further in opposite direction, the magnetization will
decrease to zero at H = He' coercivity, and reach the saturated magnetization
in the opposite direction. The parameters He' H Hss and M r , which are generally
used to characterize the behavior of a hysteresis loop, depend strongly on the
anisotropy field and temperature. They depend also on the field sweep rate,
which governs the time for the system to relax. In Fig. 5.7, we show a typical
hysteresis loop obtained by simulation, which will be discussed in more detail
in Section 5. 2. 6. The magnetization curve corresponding to the
thermodynamic equilibrium states given by Langevin function is also shown in
Fig. 5.7.

1.0
\
Hs
H
0.5

::i
:i 0.0
~

--{).5
-0.5

-1.0~""'
-1.0 ,.L...,...-_"'"
-0.1

Figure 5.7 A typical hysteresis loop, in which He' Hs ' and Mr are indicated.

5.2.6 Numerical Simulation Results


5.2.6.1
5.2.6. 1 Exponential Magnetic Relaxation
We have discussed several common phenomena observed experimentally,
160 X.X.Zhang

which are all caused by the thermally activated relaxation processes. Since it
is very difficult to have a system composed of identical magnetic particles, we
will use the numerical method to simulate the phenomena to gain a deeper
understanding of the magnetic relaxation. Suppose that a system is composed
of aligned, mono-sized, non-interacting magnetic particles whose moments are
aligned in the same direction by a magnetic field at a temperature much lower
than the blocking temperature. If the magnetic field is suddenly removed, what
will happen? Since the magnetic moment for the equilibrium state at H = 0 is
zero, the magnetic moment will relax to its equilibrium. The magnetic
relaxation will be exponential in nature for the single barrier character of the
system as (Tejada et a!. al. , 1993a)

MCt)
M( t) = Moexp( - It). (5. 10)

The anisotropy constant is chosen to be 1. 0 x 10 J/m J/ m , and the diameter of


6 3

the particles to be d. In Fig. 5.8,


5. 8, the relaxation curves are shown for particles
with different diameters at CD 300 K and ~ 400 K. It is clearly seen that when
the particle diameter is larger than 7. 2 nm, the information can be safely kept
for hundreds of years at 300 K, if these particles are used for recording
information.
It is clearly seen from Fig. 5. 8 that, just by changing the particle size from
6 nm to 7 nm, the relaxation time will wi II change for more than 10 orders in
magnitude (Fig. 5. 8a). We can roughly estimate the maximum density for the
magnetic recording if the particles, with an anisotropy constant just twice that
for hcp Co (Craik, 1995), are used as recording media. To guarantee the
stability we use the data obtained at 400 K. It is found that for particle size
larger than 7. 8 nm, recorded information can be kept for more than 100 years.
To reduce the dipolar-dipolar interactions, the particles are supposed to be put
in a lattice of 10 nm, then the density will be 6. / in = 6.
6.44 x 10 12 lin =
6.44 x G/ in 2 .
X 10 3 G/in
Thus, from the recording density point of view, we still have a long way to go
limited by the thermal activation. Of course, if
to reach the maximum density Iimited
particles with a high anisotropy are used, the size can be reduced further; and
consequently the density can be much higher.

5.2.6.
5.2. 6. 2 ZFC and FC Magnetization Curves

Since the ZFC-FC magnetization curve depends strongly on the magnetic


relaxation, the values of the magnetization at different temperatures are just an
accumulated effect of the relaxation processes during the time for the
temperature T having increased to the particular value. The differential
equation describing the dynamic behavior of a set of identical and oriented
particles is given by

dMCt) =_ [M( t) - M J/. (5.11)


dt ~
Magnetic Relaxation and Quantum Tunneling of Magnetization 161

7.2 nm

:i
~ 0.6
Q
C
~ 0.4

0.2

o
\0-11 10
10- \00 10 1 102 103 104 10
\055
Time (year)
(a)

8.0nm

:i
~ 0.6
Q
C
~ 0.4

0.2

o
10-5 10-4 10-3 10-2 10-
\0-11 10
\00 10 1 102 103 104 \05
Time (year)
(b)

Figure 5.8 Simulated magnetic relaxation for different sizes of single domain particles.
(a) 300K;
300 K; (b) 400
400K.
K.

If the temperature is changed at a constant sweep rate, {3,f3, the magnetization


value of the system is related to the relaxation rate by the expression (Tejada
eta!.,1997b):

dM dM/~ =_
= dM/Ql M (H)]r<H)
[M(H) - M
[MCH) CH)]r(H) (5.12)
C5.12)
dT dt dT eq f3
{3
where M eq is determined by the Curie law (orCor Langevin function). Since the
relaxation rate is determined by the anisotropy-energy-barrier (Eq.
CEq. (5.
C5. 1)), in
principle, the energy barrier distribution can be extracted from the ZFC-FC
data. To obtain the intrinsic energy-barrier distribution, the applied field
should be very small in comparison with the anisotropy field H K , because the
energy barrier is field-dependent as given by (Stoner
CStoner and Wohlfarth, 1948)
162 X. X. Zhang

(5. 13)

where H K = 2K ul u /M
M s is the anisotropy field, and M s is the magnetization for
single domain particles.
Now let us examine how the magnetic field affects the behavior of ZFC
and FC curves, by using the numerical simulation. In the following we will use
the parameters for Co nanoparticles of 4nm in diameter: K u = = 5. 3 x 10 5 J/m
JI m3 ,
M ss = 1446 emul
emu/cm cm (Craik, 1995), v = 10 SS-l
3 10
-1 and the temperature sweep

rate is 1/60 K/s. The ZFC and FC curves with different applied magnetic
fields are shown in Fig. 5. 9. A close examination of the ZFC curves shows that
the blocking temperature T B shifts to lower temperatures with increasing the
appl ied magnetic field, which is accounted for the reduction in the barrier
applied
height (Eq. (5. 13. The most intriguing feature in Fig. 5.9 is the peak in the
FC curves, which appears at a temperature just slightly lower than T BB. We
bel ieve that the peak in the FC curve is due to the"
believe quench" effect when the
the "quench"
sample is cooled down. When the temperature is just sl ightly below T B' the
magnetizations of the particles are fixed in their directions ("up" or "down"),
similar to an antiferromagnetic ordering at the Neel temperature. If the
temperature deceases infinitely slowly, the FC curve should follow the Curie
law. It is also seen that with increasing the magnetic field the peak becomes
smaller and finally disappears. This is because, with increasing the field, the
relaxation rate r becomes much larger due to the reduction in U, and the
system can reach a state much nearer the equilibrium. Another feature is that
above the blocking temperature, the ZFC and FC magnetization data follow
the Curie law at a low field (H = = 0.001 H K and 0.01 H K ) , but show a non-
Curie law behavior at higher fields. If 1/M 1I M is plotted as a function of
temperature, one will find that it follows the Curie-Weis law in high
temperature range with a significant negative e value, being similar to the
antiferromagnetic ordering. As we know, there is no interaction between the
particles, and thus the field induced effect would lead to confusion. This is why
appl ied field should be very small.
the applied
Since the ZFC-FC behavior is due to the relaxation, the sweep rate of
temperature should also play an important role. In Fig. 5. 10, we plot the ZFC-
FC magnetization for H = = O. 001 H K with the temperature sweep rate of
1/60 K/
KI sand 1/300 K/s
KI s to compare the behavior of the ZFC and FC curves. It
is evident that with increasing the time for the system to relax (i. e., low
sweep rate) the blocking temperature becomes smaller and the peak in the FC
curves disappears. Therefore, when the ZFC-FC data are used to extract the
energy barrier distribution, great care should be taken to choose a proper
applied magnetic field and the temperature sweep rate. If the temperature
sweep rate is too big the peak will appear in the FC curve due to the "quench"
"quenCh"
effect. These kinds of peeks may be confused with some antiferromagnetic
Magnetic Relaxation and Quantum Tunneling of Magnetization 163

Figure 5.9 The simulated ZFC and FC magnetization data with different applied fields.

0.06 0.08
0
~
0.06 FC ~ 11300 K/s
1/300

::;{ ::i 0.04


~

~ ~
0.02
ZFC ZFC
0.00 0.00
o0 L----0550:O-0-----'1C::O-:O-
100 O-----'1-'0
:0-
150
700"---:0- 0'--~300
50 ----=2200 2
250
)=c:_
300 0'----0
0 500 -----'IC::O:O-
5:0- 100 O-----'1-'0
5-:0-
150
700"---:0- 0,--7:00
0 ----=2200 2
250
)=c:_
3
300
T(K) T(K)
Figure 5. 10 ZFC-FC magnetization curves simulated with different field sweep rates.

ordering phenomena.
To observe the "quench" effect experimentally, the ZFC-FC
measurements have been performed on the Mn12 single crystals. Shown in
Fig.5. 11 is the temperature dependent magnetization measured in the
following procedures. The ZFC curve was obtained as described above (step
1), the FC-cooling curve (step 2) was obtained by cooling the sample with the
same temperature sweep rate (0. 1 K/ min) after finishing the ZFC curve.
Then, the FC-warming curve (step 3) was obtained by warming again the
sample to 4 K with the same temperature sweep rate. It is clearly seen that
the values of FC-warming magnetization are larger than the values for C-
cooling curve, particularly in the vicinity of the blocking temperature.
164 X. X. Zhang

Therefore, the "quench" effect really exists and is very difficult to be


removed.
removed,
0,0045
0.0045

0,0040
0.0040

0,0035
0.0035

0.0030
0,0030
~
S
E 0,0025
E 0.0025
~
~
::;; 0,0020
0.0020 H=2X lO-sHK
0,0015 ~
- ZFC-warming
fr-
0.0015
- FC-cooling
-0---
0-

0.0010
0,0010 - - FC-warming

0.0005
0,0005
1.5 2,0
2.0 2.5 3.0 3,5
3.5 4,0
4.0
T(K)

Figure 5. 11 The temperature dependent magnetization measured form Mn'2 in three


steps to show the"
the "quench"
quench" effect.
effect,

5.2.6.3
5.2.6. 3 Magnetic Hysteresis Loops
In the last section, the physical picture of the hysteresis loop is briefly given,
given.
Now let us use a simple model to simulate numerically the behavior of
hysteresis loops for a collection of non-interacting, single-domain particles,
particles.
The dynamic behavior of the magnetization in a magnetic field can be
described by the differential equation given below (Tejada et al.ai, , 1997b)

dMCH)
dM(H) = dMjdH
dM/dH =_ [M(H)
[MCH) _ M CH)] r(H)
(H)] C5. 14)
(5,
dH dt dt eq a
0'

and the field dependent relaxation rate given by

r(H) = U(H)
UCH)
vexp ( - ~ (5,
C5. 15)

where U(H)
UCH) is given by Eq. Eq, (5,13),
C5.13), M eq is the thermodynamic equilibrium
state described by the Langevin function, and a 0' is a the field sweep rate.
By using the parameters of Co (K CK uu = 5.3
5,3 X 10 5 J/m 3 , M s = 1,446 emul
cm 3 ) , v = 10 10 S-l a = 3.
S-1 and 0' 3, 3 X 10- 4 Tis, we calculated the hysteresis loops
at different temperatures for 4 nm Co nanoparticles.
nanoparticles, Several hysteresis loops
are shown in Fig.Fig, 5.
5, 12a.
12a, It is evident and expected that the hysteresis loops
become wider (or Cor the coercive field becomes larger) with decreasing
temperature due to the exponentially reduced relaxation rate (Eq, C5. 1.
CEq. (5. 1, To
have a close look at the temperature dependence of the coercive field, we
extracted the coercive field value from the hysteresis loops, loops. Shown in
Magnetic Relaxation and Quantum Tunneling of Magnetization 165

,
1.0
f !f
0.5

;{
~Vl

~
~ 0.0 1
f---
-- - -- -
-- - - ' --- - - -
I--
~

_25K
--25K
~.5
-0.5
--=-
--:>-35K
35 K

-1.0
JJ J --45K
-+- 55K
I
-0.4 ~.2
-0.2 0.0 0.2 0.4
H(T)
(a)

0.6 a =3.333 X 110~4


a=3.333X 0~4
o a
a =1.000 1O~5
= 1.000 X 105
X 0' =1.000 X 10-6
a=1.000XIO-u
0.4
E
f-
~
,.,..,0
:J:::u
.....
0.2

0.0

0 2 4 6 8
r (K I/2
1"2 (K
T 1i2 1.2))

(b)

Figure 5.12
5. 12 (a) hysteresis loops at different temperatures, (b) the temperature
dependence of coercive field for three different field sweep rates.

Fig.5. 12b are the coercive fields versus the temperature for three different
field sweep rates.
It is evident that the coercive field decreases with decreasing the field
sweep rate, which is due to the fact that the system has more time to relax.
The most important feature in Fig. 5. 9b is that the temperature dependence of
the coercive field can be well described by

(5. 16)

where the constant A depends on the sweep rate of the field, i. e. on the
measuring time of the equipment. It should therefore be related to the blocking
( T) = H (0) (1 - J T / T B) was
B Actually the relation He (T)
temperature, T B'
reported previously (Kneller and Luborsky, 1963; Xiao and Chien, 1987).
The relation we find here using the numerical simulation is in agreement with
the previous result.
166 X. X. Zhang

5. 3 Magnetic Relaxation in Particle Systems with Size


Distribution

5.3.1 Introduction

We have discussed the behavior of a collection of identical particles, from


which we understand some fundamental problems related to the magnetic
relaxation. Although (magnetic) particles with a very narrow size-distribution
have been fabricated recently by several groups (Sun and Murray, 1999; Sun
et al. , 2000; Wang et al. , 2000), it is still very difficult to make identical
particles. Even for a collection of the identical particles, the random alignment
al ignment
of their easy axes will cause a distribution in energy barrier, consequently, the
relaxation will change to non-exponential dependence. Another fact that should
be noted is that the interactions between the particles also cause the energy
barrier distribution. It has been found that the time dependent magnetization
usually follows the logarithmic law in many different magnetic materials (Coey
et al. , 1981; Binder and Young, 1986; Givord et aI., 1987; Labarta et al. ,
1993; Tejada, et aI., 1993b; 0' Shea and Perera, 1994; Vincent et al. ,
1994; Zhang et al. , 1995, 1996; Ibrahim et aI., 1995;). Theoretically, this
well-known law of time-dependent-magnetization originated from a flat-topped
(or constant) distribution of energy barrier (Chikazumi., 1964; Ma, 1980;
Aharoni. , 1992). Interestingly, the recent Monte Carlo simulation studies on
several typical distribution functions (nonsingular) show that all the energy
distributions lead to the universal relaxation behavior-logarithmic relaxation
(Gonzalez-Miranda and Tejada, 1994), although some other relaxation laws
have been suggested (Skomski and Christoph, 1989; Aharoni, 1992).
Actually, logarithmic magnetic relaxation is commonly observed in magnetic
systems and superconductors (Yeshurun et al. , 1996).
So far, exponential magnetic relaxation, M ( t) = M (0) e - n ,
(Eq. (5.10))
(5.10 has only been observed in TbFe03 single crystal (Zhang et al. ,
1994) and Mn12 magnetic molecular crystal (Friedman et al., aI., 1996). The
universal barrier in Mn12 can easily be understood as described in Section 2. 2.
Whereas in the TbFe03singie
TbFe03 single crystal, there are several possible origins for the
universal barrier in the bulk sample, but which one dominates is not clear yet
(Zhang et al.,
aI., 1994). To have a clear picture of the exponential magnetic
relaxation, we show, in Fig. 5. 13, the relaxation data obtained from MnMn1212and
and
TbFe03' The relaxation measurements were carried out as follows. For Mn12 Mn12',
the sample was magnetized (saturated) with a magnetic field of 3 T, and then
Magnetic Relaxation and Quantum Tunneling of Magnetization 167

the field was switched off and the magnetization was measured as a function of
TbFeD 3 single crystal, the sample was
time for several hours. Whereas, for TbFe03
De field, then the field
cooled from 300 K to the measuring temperature in a 20 Oe
was changed to - 200 Oe De and the time-dependent magnetization was
measured.
-4.800
-5.000 -2.65
-5.200 -2.70
E -5.400 =
::l
a
E
~-5.600 ~ -2.75
"c'
'c' -5.800 -=
.s -2.80
- -6.000
--{).OOO
--{).200
-6.200 -2.85
--{).40oL....o...~~~~~_~~--'-~_
-6.400
~~'-----:'c~~~,,-;:--'---:::~~--,-,,'= -2.90!:-~--:7::-::-~=':::::-~::-::'::"::--~"""
-2.90!:-~--,-"'=-~~:::-~::-::,::,::--~.,..,,
o 1000 2000 3000 4000 0 1000 2000 3000 4000
500 1500 2500 3500 500 1500 2500 3500
Time (s)
(5) Time (s)
(5)

---.-- 0.070
.-.-.- ......-. ..
.-.-.-.-.-.....~

~
0.006 0.068

"\ ~\
0.066
S'
S-
a
E 0.004
1 0.064
~ ~ 0.062
~ ~ 0.060
0.058
0.002
0.056
0.054L-_--'-_
0.054l-_~ _ _ _~~ __ _~~ __
_~~__
4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 4 5 6 7 8
In (time) In (time)
(a) (b)

5. 13 The magnetic relaxation data obtained from Mnl2


Figure S. Mn12 single crystal (a) and
TbFe03 single crystal (b) are plotted in the semi-logarithmic scale.
scale, i.e
i. e.., In(M) versus T
(upper) and M versus In (t) (bottom). The relaxation data were taken at 2. 6 K and 2 K
for Mnl2
Mn12 and TbFe03 respectively.

It is evident that the magnetic relaxation data can be well fitted to the
exponential magnetic relaxation rather than the logarithmic dependence.
Actually, they are the only systems showing the exponential magnetic
relaxation.

5.3.2 Logarithmic Magnetic Relaxation


In this section we will give aa brief derivation of the logarithmic relaxation law
and a non-singular
from Eq. (5. 10) and non-singular energy barrier distribution in order to give
give
a clear physical picture
picture of magnetic
magnetic relaxation
relaxation process. Furthermore, wewe may
may
gain a deeper
deeper understanding
understanding of the relationship
relationship between the magnetic viscosity
168 X. X. Zhang

and the energy barrier distribution. For a system with an energy barrier
distribution f( U), the relaxation will be given by

M(0) dUf(
M( t) = M(O) f
U) e- rtn
d Uf ( U)e- (5.17)

f
where dUf( U) = 1. The time derivative of Eq. (5. 17) could be written as
(Tejada et al.
a!. , 1993a)
~
~

dM =
dt
dT =-- M(O) k Tf dxe -Xf[k ss Tln( vtl x) ]
-t-
T (5.18)
(5. 18)
o

where the relation of x = nand


It and dU = = -- (k s Til) dl were used. For a
nonsingular distribution function, the major contribution to the integral comes
from x - 1. This means that only the metastable states with a 1-1 lifetime
comparable to the observation time contribute to the integral. For a typical
relaxation measurement t -1- 1 - 10 4 s, that is, vt - 10 10 - 10 14 . At x - 1, it
gives In( vt)>>ln(x), so that Eq. (5. 18)can be approximated with accuracy as

dM ks T
dt
dT =--
= M(O) -t-f[k s Tln( vt)]. (5.19)

Integrating Eq. (5. 19)9ives


y( t)

f
y(t)

M(t) = MCto)[l-
M( t) = MCt o ) [1 - dy'f(y') ] (5.20)
o
= V-I _10- 10 S,
where to = s, M (0) is replaced by M ((to)
M(O) to) for a proper treatment of
small times,
yet)
y( t) = k s Tln(vt)
Tln( vt) (5.21)

A very important observation follows from Eqs. (5. 20) and (5. 21 ): M
depends on t through Eq. (5. 20). This means that if a logarithmic relaxation is
observed due to the energy distribution, the coefficient in front of 10gCt)
log( t) must
be proportional to temperature T. The commonly observed 10{Jarithmic 10(;Jarithmic
relaxation can be easily understood from Eq. (5. 20). Let <U> be the average
energy barrier which determines the blocking temperature measured in ZFC
curve, <U> -30k s T s . At T T s ' the integral in Eq (5.20) can be
approximated by the first term of the expansion in series on y I W>

M(t) = MCto)[l- ~~~lnCtlto) l (5.22)

Eq. (5. 22) is valid for the low-temperature relaxation over the small barriers,
and the relaxation part is small in comparison to the total moment. It should be
also expected that this expression is invalid for a large t when the temperature
Magnetic Relaxation and Quantum Tunneling of Magnetization 169

is comparable to the blocking temperature. The value of magnetic viscosity,

S = ksT (5.23)
<U>
is proportional to the temperature, which determines the speed of the
relaxation. When the temperature approaches absolute zero, S becomes
zero.
Everything that we have discussed above is related to over barrier
transition caused by the thermal activation processes. Both the relaxation rate
r and the magnetic viscosity will vanish, when T approaches to zero. This
can be easily understood as that when the temperature becomes zero, all the
thermal processes will be completely ruled out.
Another interesting feature observed from Eq. (5. 22) is that the time
dependent magnetization M(
M ( t) is a function of T In(
In ( vt). This means that the
magnetic relaxation data measured at different temperatures below the
blocking temperature T s ' scale with T In( vt) (Tejada et al.
a!. , 1993a; Labarta
et al., a!., 1994). Although v is weakly temperature
a!., 1993; Vincent et al.,
dependent and there is a distribution of v values in complex systems, the T In
vt) plot remains good, to a first-order approximation.
((vt)
In Fig. 5. 14a, we present the magnetization data obtained on y-Fe2 0 3
nanoparticles, with an average size of 10 nm, diluted in water to form a
magnetic fluid (Ziolo et a!.,
aI., 1992; Zhang et a!.,
aI., 1995). Since the size-
distribution of particles, the ZFC curve measured with an applied
appl ied magnetic
field of 0.01
O. 01 T shows a broad peak. The relaxation measurements were
performed as follows. The sample was cooled from 250 K to a target
temperature in a O. 01 T field (FC), then, the field was changed to -0.01
- O. 01 T.
The time dependent magnetization was measured in few hours. In Fig. 5. 14b
we plot the magnetization as a function of In( t) for different temperatures from
1. 8 to 35 K. It is obvious that the magnetic relaxation curves are logarithmic in
nature. The values of the magnetic viscosity were extracted by fitting the
magnetization data to Eq. (5. 22). In Fig. 5. 14c, the values of the magnetic
viscosity are plotted versus the temperature. A linear dependence of magnetic
viscosity S on the temperature is observed in the whole temperature range,
indicating a pure thermally activated relaxation process. We also plotted in
Fig.5. 14d the magnetic relaxation data obtained at different temperatures as a
function of a single variable T In ( vt), (the scaling
scal ing law). The important
feature in this figure is that all the data collapse into a master-curve, the
scaling law, indicating again that the data follow a logarithmic law and are of
thermal origin.
We know that an energy barrier distribution leads to a logarithmic
170 X.X.Zhang
X. X. Zhang

.!.=;~!::;!....!;!;... :;~:-
.!.::==;!=;~!...!;!;~~;:.
0.12 0.10 :===:=;-~:-
:===:=;-~:_ ..: ...-.-
.... .. _e_ _

;;~ :-:~-=;.:;;;[;~;=:-
:~;;.~~;;[;;==.-
0.10 O.OS
"? 0.08
_--
---
::J
:::l ::J
E 0.08
O.OS !.E
~
0.06 ._-.-.-.-.-
.-.-.-.-.-
'" 0.06
~
'lJ

OJJ6 ~ 0.04 '.==:=:=:::::::::=-


'.==:=:~::::::::::--
0.04 ZFC 0.02
0.02 .-.-.-._~
.-.-.-.-~..._ _ _-_
0.02
O.OOu,--~-~----=-----=----o-
O. 00l..L..---=----'c------=-~---7-
o 50 ISO 200 250
100 150 4 5 6
5 6 77 S
8 9
T(K) In (t)
(I)
(a)
(al (b)

40 0.12
0.10 ~,
~~
~
30
'c
'2
~
:J "? 0.08
E 0.08 '
" "
-e~0 20 2~ 0.06 "
~""'
v;
'" 10
'~ 0.04
0.04
0.02
0.02
"
"
~,
"
20 30 40 o 400 800
SOO 1200
T(K) Tln(vt)
(c) (d)

Figure 5.14
5. 14 (a) ZFC-FC magnetization with a field of 0
O. 01 T
T. (b) Magnetization
relaxation at different temperatures from 1. 8 to 35 K. (c) The values of the magnetic
viscosity extracted from (b) as a function of temperature. (d) Magnetization relaxation
data shown are plotted as a single variable T In (vi), and v = 10 '0 s- 1.

magnetic relaxation rather than the exponential relaxation. The interesting


point is how an energy distribution affects the ZFC-FC curve. Since it is very
difficult to have a system with well controlled distribution, we use numerical
simulation to examine the effect of the energy barrier distribution by just adding
a Gaussian distribution to a particle's radius, i. e. , R o
0 - Li~Ro~Ro
- 1:. ~ R0 ~ R0 + 1:..
Li. To
compare the effect of the distribution, we will use the parameters for 4 nm Co
particles used in Section 2. 5. We have used the following values to change the
width of the distribution, 1:.
Li = 1% (0.2 A), 5% (1 A), 10% (2 A), 20%
(4 A) and 30% (6 A). Shown in Fig. 5. 15 are the ZFC-FC curves simulated
with different distribution ranges. It is seen that quite a narrow distribution can
significantly affect the behavior of ZFC, i. e. , the ZFC increases much faster
at low temperatures (or the sharp increase in ZFC near the blocking
temperature shown in Fig. 5. 7 changes to much smoother one). The behavior
shown in Fig. 5. 15e, f has been commonly observed experimentally in particle
systems. We now understand that the characteristic behavior of ZFC-FC (Fig.
5. 15e, f) is due to the distribution of the energy barrier. The wider the
distribution, the broader the peak.
Magnetic Relaxation and Quantum Tunneling of Magnetization 171

0.06 ,1=0 0.06 ,1=I%Ro

~
.,,; 0.04 0.04
"'""
~
;:;::;
~
""" 0.02 0.02

0.00 0.00
o0 50 100 150 200 o
0 50 100 150 200
(a) (b)

0.06 ,1=5%Ro 0.06 ,1=10%Ru


-;'I, ~
.,,; 0.04
::{ 0.04 0.04
"""
~

~ 0.02
~ 0.02

o.oo"----;';_--;-;;-----;-:~~,----;:;~
0.00 0.00
o0 50 100 150 200
(c)
(e)

0.06 0.06
,1=30%R o
0.05
~~ 0.04
.,,; 0.04
"""
~ 0.03
~
~ 0.02 0.02
0.01
0.00 O.OO"----;';,-------;-;;-----;-:~~,----;:;~
0,00
0o 50 100 150 200 o 50 100 150 200
T(K) T(K)
(e) (f)

Figure 5. 15 The simulated ZFC-FC magnetization data for particles with a Gaussian
distribution in radius.

5.4 Quantum Tunneling of Magnetization

5.
5.4.
4. 1 Introduction

The facts that the relaxation rate decreases with decreasing the temperature
and the relaxation vanishes at the absolute zero temperature, seem to tell us
that the recording density can reach very high levels, provided that the
recording media is kept at very low temperatures. Unfortunately, this is
proved not true by the observation of macroscopic quantum tunneling of
magnetization (QTM), which was predicted by Chudnovsky et al. in 1988
172 X. X. Zhang

theoretically (Chudnovsky and Gunther, 1988) and since then QTM has been a
focus in condensed matter physics both in theory and experiments (Barbara
and Chudnovsky, 1990; Stamp, 1991; Zhang et al., a!., 1992; Tejada et al.
a!. ,
a!., 1995; Chudnovsky and Tejada, 1998).
1993a,b, 1996, 1997a; Gider et al.,

5.4.2 Physics Related to QTM


The physical picture of QTM is that at very low temperatures the transition of
the magnetic moment will be dominated by the under-barrier quantum
mechanical tunneling processes rather than through the thermal activated over-
barrier processes. That is, the relaxation rate will be given by a constant
value rI Q '-"- vexp
vexp(( - U / k B T c) --
'" exp
exp(( - B)
B),, when the temperature is below the
crossover temperature T c' where the constant B is the WKB exponent
(Chudnovsky and Tejada, 1998). For v'" v -- 10 10
10
-- 1011 S-I, at k B
lOll S-l, B T U/30,
thermal transitions are frozen out. To observe the tunneling effect of
magnetization in the relaxation measurement carried out in a typical time
window--10
window'" 10 4 s, the following condition should be satisfied:

kB T < U/30 :s;;;


~ k B Tc (5.24)

where U=KV(1-H/H a )2, i.e., Eq.(5.13).


Eq. (5.13).
The next question are: At which temperature can the quantum tunneling
dominated magnetic relaxation be observed and what is the tunneling rate of
the magnetic moments? Here, we will give a physical picture of QTM in
magnetic nanoparticles and answer the above questions with a brief
introduction to several theoretical models of magnetic tunnel ing (Chudnovsky
and Tejada, 1998). For tunneling of the magnetic moment in ferromagnetic
nanoparticles, four models were described in the book Macroscopic quantum
tunneling of the magnetic moment (Chudnovsky and Tejada, 1998).
The first model describes tunneling of the magnetic moment in a single-
domain particle with volume V, magnetization M o and total anisotropy energy

E = k.l M~ - k ll M;
M~ = V(K.l cos 2 e
e-- K II sin 2 ecos 2 cf
</ (5.25)

where K K II and K.l are parallel and transverse anisotropy constants


respectively. For K II > 0 and K.l > 0, Eq. (5. 25) indicates the x-y easy
plane and the x easy axis in the plane. It is evident that the energy given by
Eq. (5. 25) has two degenerate minima corresponding to M pointing to positive
and negative X directions, i.e., cf>=0,
i. e., </> = 0, "IT e = "IT/2. If M flips
IT at e=lT/2. fl ips in the easy
plane (x-y plane), the barrier separating the two minima is U = VK II ;
whereas the barrier will be U = = V (K II + K.l)' for M to flip through the x-z
plane. The theoretical calculation gives the tunneling rate as

Ir~1
~ I cos(SlT)
cos ( S"IT) I Wo
w oexp - 2 S In [ ( 1 +
ex P {{-2Sln[(1 ~~~~)1/2
~ ) + (~
1/2 (~~)1/2J}
I: ) 1/2 ] ) (5.26)
Magnetic Relaxation and Quantum Tunneling of Magnetization 173

with

Wo = ~[KIIII (K II + KK..L)]1/2
= :J:[K J)JI/2 (5.27)

where S = M o V V/I((liy) ...., 10 10 -


hy) is the total spin in the particle, and y ((....,
1011 S -1) is the ferromagnetic resonance (FMR) frequency. It should be noted
K..L *0, i. e. , when M x does not commute with
that tunneling happens only for K..I-
the Hamiltonian (Chudnovsky and Tejada, 1998). For an integer S, Eq.
(5.26) gives

(4;: )]
= Wo
rI '" woex p [ - Sln[ (4:: ) ] = (4~~1
(4~~1 )) s (5.28)

K..L K II ', and


at K..I-

rI '" woexp [ - 2Sln[ (~~


(~ ~ ) 1/2 ] (5.29)

K..L K II' The latter case is more interesting for tunneling of large spin and
at K..I-
can be appl ied to Tb and Dy whose anisotropy constants are K II ~ 106 erg/ergl cm 3
and K 8
erg I cm . Using these values and Eqs. (5. 27) and (5. 29), one
K II ~ 10 erg/ 3

finds that wo~10 11Hz and the tunneling rate 1~1011exp(-0.


r~1011exp(-0. 2S) Hz.
It is understood that the tunneling rate should be equal to the thermal
transition rate, rITT cc K II V /IT)
ex: exp ( - K T) at the crossover temperature T c
(Chudnovsky and Gunther, 1988). One, therefore, finds that the quantum
tunneling effect dominates the spin flipping below
T c = IJBH II /In(4H II /H..L) (5.30)

K..L K II and below


for K..I-

T cc = ~ IJIJB(H
B ( H II H..L) 1/2
II H..I-) (5.31)

K..L K II (Chudnovsky and Tejada, 1998). Here we have introduced the


for K..I-
anisotropy field H 11(..1-)
11<..Ll == 2K 11(-lJ
11<..Ll I/ Mo.
Mo
Model 2 is about the tunneling in a single-domain particle whose uniaxial
anisotropy axis is along the z-axis In addition a transverse field H is applied in
the x -axis. In this case, the energy of the particle is (Chudnovsky and
Tejada, 1998)
2
2
H2 M ))
= VV ( Ksin 2
E = e- HMosinecos +~ . (5.32)

For the applied field H< H<H 2K I M o (the anisotropy field), energy E given
=2K/M
H.a =
by Eq. (5. 32) has two minima (E min = 0) for e = eo
(Em;n eo and e = IT
TT - eo
eo at = 0,
eo
where eo == HHI/ HH.a' The energy barrier between the two minima is U == KV (1 -
= KVE 22 .. The calculated tunneling exponent B is given by
H/H a )2 =
HIH.)2

2Sln( ~a)
B = 2Sln(%) (5.33)
174 X.X.Zhang
X. X. Zhang

for HH a , and
= 4SE 3/2
B = 3 2
/ (5.34)
for H-Ha(or E ........ O).
H---+Ha(or E-O).
In the third model, we discuss magnetic tunneling in the particle that is
used for Modell and now is put into a magnetic field being applied along the
negative x axis, opposite to the magnetic moment of the particle. The energy
of the particles becomes CChudnovsky
(Chudnovsky and Tejada, 1998)
E = k ~ M~ - k II M~ + HM x
= V(K~ cos 2 e - K II sin 2 ecos 2 <f>
+ HMosinecos).
HMosinecos<f. (5.35)
In this case, the energy (Eq. (5. 35 has two not-equivalent minima, i. e. , M
pointing to positive and negative x directions correspond respectively to the
metastable state and absolute minimum of energy. The energy barrier
separating these two minima is U = K II V ( 1 - H / H II )2 = K II VE 2 . For the
tunnel ing, i. e., K
most favorable situation for tunneling, ~
K~ II and H-
K II
H ---+ H II'
II , the
tunneling rate is

r ""
""-' W HQ ( H) exp [ -
wHa(H)exp K:
38 S (KK: )1/2
112 3 2 ]
E
E 3/2
/ (5.36)

where Q
a (H) is a dimensionless function oscillating in field (Chudnovsky and
Tejada, 1998), and

WH -- 2 Y[ EK II K ~ J
- M
2y[ 1/ 2
J1/2
. (5.37)
o
The fourth model is more generic and the easiest to implement experimentally,
in which the magnetic field is appl
applied eH (#0, 1TTI/2
ied at some angle eH(#O, /2 etc.)
etc. ) to the
anisotropy axis. This problem does not posses any symmetry and therefore is
more difficult mathematically. Interested readers can conrefer
con refer to Chudnovsky
and Tejada (1998). Due to the random alignment of the anisotropy axes and
size distribution of the nanoparticles in the samples used for experiments, it
may be difficult to compare the theories with experimental results directly. But
the crossover temperature and the tunneling
tunnel ing rate can be roughly estimated by
using these models (Eq. (5. 26) - (5.37 and the materials properties.
Tunneling of Neel vectors in anitiferromagnetic particles are also very
interesting (Barbara and Chudnovsky, 1990) and have been observed
experimentally (Awschalom et al.,aI., 1992; Tejada et al., 1997a). It is well
known that the anisotropy in the antiferromagnets is greatly enhanced by the
exchange coupling (Chudnovsky and Tejada, 1998). Therefore, the tunneling
effect of Neel vectors is much stronger than the tunneling of the magnetic
moment in ferromagnets, because the crossover temperature scales with the
anisotropy field (Eq. (5. 31)
31 . Since the Neel vector does not couple to any
experimental field (magnetization M = = 0), the antiferromagnetic tunneling can
be detected by magnetic measurement only if there is a small magnetic
moment resulting from uncompensation of the sub sublattices.
lattices . Fortunately, there
Magnetic Relaxation and Quantum Tunneling of Magnetization 175

are always uncompensated surface spins in the nanometer-sized


antiferromagnetic particles due to the different sizes of the sublattices or the
irregular shape of the particles. For the tetragonal anisotropy, the tunneling
rate is
is. calculated as (Chudnovsky and Tejada, 1998)

2V ( 2 K ) 1/2 ]
I ""I
""'I cos (SlT) 1 woexp [ - hy 2X-l K II + m K~
1 <.Vaexp K: (5.38)

with

_y
_ Y 2K I IK- l ) 1/2
( m 2 +2X-l K-l (5.39)
<.Va --
Wo

where S is the net spin of the particle; m is the sub-lattice magnetization; X-l
is the transverse susceptibility. In almost compensated antiferromagnetic
particles, the tunneling effect dominated the flipping of Neel vectors below
(5.40)
It is evident that the crossover temperature in antiferromagnets should be
much larger than that in ferromagnets because of the fact that Hex is generally
much larger than H-l'
H -l' Here we just give a brief review of the tunneling rate
and crossover temperature in the ferromagnetic and antiferromagnetic
particles, one can find the detailed theories in the book Macroscopic quantum
tunneling of the magnetic moment (Chudnovsky and Tejada, 1998). In the
follow section, we will give an example of experimental observation of QTM.

5.4.3 Observation of QTM in Systems of Nanostructured Materials


Quantum tunneling of magnetization has been observed experimentally in
different systems by measuring the magnetic relaxation at different
temperatures (Zhang et al. , 1992,1996; Paulsen et aI.,al. , 1992; Tejada at al. ,
1993a, 1993b, 1993c, 1996, 1997a; Arnaudas et aI., 1993; Ibrahim et aI., al. ,
1995; Chudnovsky and Tejada, 1998). From the relaxation data, the
relaxation rate (n(n for a system with single barrier or magnetic viscosity (S)
for a system with a barrier distribution can be extracted. By analyzing the
temperature dependent relaxation rate (or magnetic viscosity), one can find
that the quantum tunneling dominates the transition of the moments below a
certain temperature-crossover temperature T c' The feature for the quantum
tunneling of magnetization is that the relaxation rate I (or viscosity S)
approaches a constant below T c .
Here we will show the experiment on the CoFe 2 04 magnetic particles in a
magnetic fluid (Tejada et al.,
aI., 1996). For the purpose of comparison, the
ZFC-FC magnetization curves, relaxation data, magnetic viscosity and M
versus T In(
In (vt)
vt) are shown in Fig. 5. 16. From the hysteresis loops measured
at different temperatures, it is known that at 5 K the anisotropy field is at least
176 X.X.Zhang

5 T (Tejada et al. , 1996). Therefore, the crossover temperature should be at


least few K, according to the discussion in the last section.
section, It is also clearly
seen from the ZFC curve that the average blocking temperature is about (or
higher than) 250 K. When the applied magnetic field is increased to O. 0 4 T, the
average blocking temperature is shifted to"-'
to""'" 160 K, in agreement with the fact
that the blocking temperature decreases with increasing the field. The
relaxation measurements were carried out by cooling the sample from 250 K to
the target temperatures in a field of O. 5 T, and then by changing the field
- 0.4 T for the relaxation measurement. All relaxation data measured at
different temperatures can be well described by the logarithmic law as shown
in Fig.5. 16b. The magnetic viscosity values for all temperatures were
extracted from fitting the data to Eq. (5. 22) , which are plotted as a function of
temperature in Fig. 5. 16c. The most interesting feature in this figure is that
below 2. 4 K, the magnetic viscosity approaches a constant, the feature for
the macroscopic quantum tunneling
tunnel ing of magnetization. In addition, in the high
temperature regime, magnetic viscosity is proportional to temperature, the
characteristic for thermal activated process. In Fig. 5. 16d, the relaxation data
are plotted as a function of Tin ( vt), i. e., the scaling law for thermal
activated relaxation. If the relaxation process is caused by quantum tunnel ing,
tunneling,
0.04
,,:100000
0.168
".co e)
Fe
<DpCOOOOOOOOOOOOQ
'OCO':'CJ
0'C,
0 0 '0
S'
OJ 0,
00 0.166
D C ,,_,.
0 S'
=>
E 0.02 .,'~o ;:::
~ .....
.,...'.
. r..i E
~ 0.164
~

...,.-..........'.
-,.
'

~
'< ~
~
~~.~......
'
""" 0.162
0.00 .....
ZFC ......'.
0.160 ,
0 50 100 150 200 250 300 3 ., 4 5 6 7 8 9
T(K) In (t)
(I)
(a) (b)
5 0.170
~",_4.2K
K
~ 4 0.165 "",
''c";;;
:::;
::J
.D
3
S'
E
E
~
~ 0.160
~ 2 ~
~
< ~4K K
V:>
Vj
0.155 c'""

0 4 5 8 10
0.150
40 80 120 160
"
T(K) nn(vt)
Tln(vf)
((c)
c) (d)

Figure 5. 16 Magnetization data obtained in different measurements on CoFe,O,


nanoparticles with diameter of 3 nm in water. (a) ZFC-FC magnetization with a 0.01O. 01 T
applied field. (b) Representative magnetization relaxation data obtained at different
temperatures from 18
1,8 to 8 K.
K (c) Magnetic viscosities extracted from (b) as a function of
temperature. (d) Magnetization relaxation data shown in (b) are plotted as a single
variable T In v= 1O 'IOs-
(vI) and v=10 '.
s-'.
Magnetic Relaxation and Quantum Tunneling
Tunnel ing of Magnetization 177

the relaxation data should not follow the scaling law. The relaxation data follow
the scaling law for T> 4 K, but deviate from the scaling law for T < 4 K,
evidence of the occurrence of the quantum tunneling of magnetization. By
comparing the data shown in Fig. 5. 14 with those in Fig. 5. 16, one can see
clearly the features of macroscopic quantum tunneling of magnetization in the
temperature dependent viscosity and M '" - Tin ( vt) plot. The absence of
quantum tunneling of magnetization in y-Fe2 0 3 nanoparticles at the
temperature down to 1. 8 K is due to the fact that their magnetic anisotropy is
almost 10 times lower than that in CoFe
CoFe204
20 4 nanoparticles, and therefore the
crossover temperature should be much lower than 2. 4 K.

5.5 Conclusions and Future Perspective

We have discussed the magnetic relaxation and related phenomena. The


numerical approach has been used to simulate the different behaviors, which
are usually used for characterization of the magnetic nanoparticles. The
physics on the macroscopic quantum tunneling of magnetization has been
discussed briefly and the experimental observation of the quantum tunneling
effect in CoFe20 44 nanoparticles is presented as an example.
Since the discovery of spin resonant quantum tunneling in the high spin
magnetic molecules, Mn12 and Fes,
FeB' much effort has been dedicated to the new
al., 1996; Hernandez et al.,
physics (Friedman et aI., aI., 1996; Thomas et al. ,
1996; Chudnovsky et aI.,
al. , 1997; Prokof' ev et aI., 1998; Kim, 1999; Zhong
et al. , 2000; Bokacheva et al. , 2000; Luis et al. , 2000; Sangregorio, 1997;
Zhang et aI., 1999; del Barco et aI., 1999). Actually this phenomenon is
closely related to the magnetic relaxation and there are still some interesting
fundamental problems that are being intensively studied (Chudnovsky and
Garanin, 2001).

References
Aharoni, A. Phys. Rev. B 46: 5,434(1992)
846: 5 ,434( 1992)
Arnaudas, J. I.,
I. , A. del Moral, C. de la Fuente and P. A. J. de Groot. Phys.
B47: 11,924(1993)
Rev. 847:
Awschalom, D. D., M. A. McCord and G. Grinstein. Phys. Rev. Lett. 65:
783(1990)
Awschalom, D. D., J. F. Smyth, G. Grinstein, D. P. DiVincenzo and D.
3 ,092( 1992)
Loss. Phys. Rev. Lett. 68: 3,092(1992)
B. and E. M. Chudnovsky. Phys. Lett. A 145: 205 (1990)
Barbara, B.and
Barbara, B., L. C. Sampaio, J. E. Wegrowe, B. A. Ratnam, A.
Marchand, C. Paulsen, M. A. Novak, J. L. Tholence, M. Uehara and D.
178 X.X.Zhang

Fruchart. J. Appl. Phys. 73: 6, 703( 1993)


6,703(1993)
Becker, R. and W. Doring. Ferromagnetismum. Springer, Berlin (1939)
Binder, K. and A.P. Young. Rev. Mod. Phys. 58: 801(1986)
Bokacheva, L. , A. D. Kent and M. A. Walters. Phys. Rev. Lett. 85: 4803
A.D.
(2000)
Chantrell, R. W. and E. P. Wohlfarth. J. Magn. Magn. Mater. 40: 1 (1983)
Chen, J. P., C. M. Sorensen, K. J. Klabunde and G. C. Hadjipanayis.
Hadj ipanayis.
Phys. Rev. B 51: 11,527(1995)
B51:
Chien, C. L. J. Appl. Phys. 69: 5 ,267( 1991)
5,267(1991)
Chikazumi, S. Physics of Magnetism. Wiley, New York (1964)
Chuang, D.S., C. A. Ballentine and R. C. O'Handley. Phys. Rev. B 49:
15,084(1994)
Chudnovsky, E. M. and L. Gunther. Phys. Rev. Lett. 60: 661 (1988); Phys.
9,455 ( 1988)
Rev. B 37: 9,455(1988)
Chudnovsky, E. M. and D. A. Garanin. Phys. Rev. Lett. 79: 4,469 ( 1997)
Chudnovsky, E. M. and J. Tejada. Macroscopic quantum tunneling of the
magnetic moment. Cambridge University Press, Cambridge ( 1998) and
references therein
Chudnovsky, E. M. and D. A. Garanin. Phys. Rev. Lett. 87: 187,203
(2001), and references therein
Coey, J. M. D., T. R. McGuire and B. Tissier. Phys. Rev. B 24: 1,261
( 1981)
Craik, D. Magnetism: principles and applications John Wiley & Sons Ltd,
p.404(1995)
Cullity, B.D.
B. D. Introduction to Magnetic Materials. Addison-Wesley, Reading,
MA (1972)
del Barco, E. , N. Vernier, J. M. Hernandez, J. Tejada, E. M. Chudnovsky,
E. Molins and G. Bellessa. Europhys. Lett. 47: 722( 1999)
Dickson, D. P. E., N. M. R. Reid, C. Hunt, J. D. Williams, M. EI-Hilo, K.
O'Grady. J. Magn. Magn. Mater. 125: 345(1993)
Djurberg, C. , P. Svedl indh, P. Nordblad, M. F. Hansen, F. B0dker and S.
M0rup. Phys. Rev. Lett. 79: 5,154(1997)
Dormann, J. L, L. Bessais and D. Fiorani. J. Phys. C 21: 2,015(1988)
Dormann, J. L., D. Fiorani and E. Trone. Adv. Chem. Phys. 98: 283
( 1997)
Fannin, P. C. A. Molina and S. W. Carlest. J. Phys. D: 0: Appl. Phys. 26:
2006( 1993)
2006(1993)
Friedman, J. R., M. R. Sarachik, J. Tejada and R. F. Ziolo. Phys. Rev.
Lett. 76: 3,830(1996)
Garcia-Otero, J. , M. Porto, J. Rivas and A. Bunde. Phys. Rev. Lett. 84:
167(2000)
Gider, S., D. D. Awschalom, T. Douglas, S. Mann and M. Chaparala.
Science 268: 77 ( 1995), and references therein
Givord, D., A. Lienard, P. Tenaud and T. Viadieu. J. Magn. & Magn.
Magnetic Relaxation and Quantum Tunneling of Magnetization 179

Mater. 67: L281( 1987)


3 ,867( 1994)
Gonzalez-Miranda, J. M. and J. Tejada. Phys. Rev. B 49: 3,867(1994)
Hernandez, J. M.,M. , X. X. Zhang, F Luis, J. Bartolome, J. Tejada and R.
Ziolo. Europhys. Lett. 35: 301 301(( 1996)
Hernandez, J. M. M.,, X. X. Zhang, F. Luis, J. Tejada, J. R. Friedman,M. P.
Sarachik and R. Ziolo. Phys. Rev. B 55: 5,858 (1997)
Ibrahim, M. M.,M. , S. Darwish and M. M. Seehra. Phys. Rev. B 51: 2,955
(1995)
Johansson, C., M. Hanson, P. V. Hendriksen, S. Morup, J. Magn. &
Magn. Mater. 122: 125(1993)
Jonsson, T. , J. Mattsson, C. Djurberg, F. A. Khan, P. Nordblad and P.
Svedlindh. Phys. Rev. Lett. 75: 4, 138( 1995)
4,138(1995)
Jonsson, T., P. Svedlindh and M. F. Hansen. Phys. Rev. Lett. 81: 3,976
(1998)
Kim, G.H. Phys. Rev. 859: B 59: 11,847(1999)
Kneller, E. F. and F. E. Luborsky. J. Appl. Phys. 34: 656 ( 1963)
Kodama, R. H., C. L. Seaman, A. E. Berkowitz and M. B. Maple. J.
Appl. Phys. 75: 5,639(1994)
5 ,639( 1994)
Labarta, A., O. Iglesias, L. Balcells and F. Badia. Phys. Rev. B 48:
10,240 (1993)
Linderoth, S. , L. Balcells, A. Labarta, J. Tejada, P. V. Hendriksen, S. A.
Sethi. J. Magn. & Magn. Mater. 124: 269(1993)
Luo, W., S. R. Nagel, T. F. Rosenbaum and R. E. Rosensweig. Phys.
Rev. Lett. 67: 2, 721( 1991)
Luis, F. , F. L. Mettes, J. Tejada, D. Gatteschi and L. J. De Jongh. Phys.
Rev. Lett. 85: 4,377(2000)
4,377 (2000)
Ma, S.K. Phys. Rev. B22: 4,484(1980)
Mamiya, H., I. Nakatani and T. Furubayashi, Phys. Rev. Lett. 80: 177
(1998)
M0rup, S. and E. Tronc.
Trone. Phys. Rev. Lett. 72: 3 ,278( 1994)
Mydoch, J. A. Spin Glasses: An Experimental Introduction. Taylor &
Francis, Washington, DC (1993)
Neel, L. C.R.
C. R. Acad. Sci. Paris 228, 664(1949a)
664 ( 1949a)
Neel L. Ann. Geophys. 5: 99( 99 (1949b)
1949b)
O'Shea, M. J. and P. Perera J. Appl. Phys. 76: 6174(1994)
Paulsen, C., L. C. Sampaio, B. Barbara, R. Tucoulu-Tachoneres, D.
Fruchart,, A. Marchand, J. L. Tholence and M. Uehara. Europhys. Lett.
Fruehart
19: 643(1992)
Prene, P. , E. Tronc,
Trone, J. P. Jolivet, J. Livage, R. Cherkaoui, M. Nogues,
J. L. Dormann and D. Fiorani. IEEE Trans. Magn. 29: 2658(1993)
Prokof'ev,
Prokof' ev, N.V.
N. V. and P.C.E.
P. C. E. Stamp. Phys. Rev. Lett. 80: 5794(1998)
5794( 1998)
Sangregorio, C. , T. Ohm, C. Paulsen, R. Sessoli and D. Gatteschi, Phy.
Rev. Lett. 78: 4645 ( 1997)
Sappey, R. , E. Vincent, M. Ocio, J. Hammann, F. Chaput, J. P. Boilot and
180 X. X. Zhang
X.X.Zhang

D. Zins. Europhys. Lett. 37: 639 ( 1997)


Shtrikman, S. and E. P. Wohlfarth. Phys. Lett. A 85: 467 (1981)
Skomski, R. and V. Christoph. Phys. Stat. Sol. (b) 156: K149(1989)
P. C. E. Phys. Rev. Lett. 66: 2802 (1991)
Stamp, P.C.E.
Stamp, P. C. E., E. M. Chudnovsky and B. Barbara, Int. J. Mod. Phys. B
6: 1355 ( 1992)
Stoner, E.C.
E. C. and E. P. Wohlfarth. Philos. Trans. London Ser. A 240: 599
(1948)
Sucksmith, W. A. and J. E. Thompson. Proc. R. Roc. London Ser. A 225:
362(1954)
362( 1954)
Sun, S. H. and C. B. Murray. J. Appl. Phys. 85: 4325 ( 1999)
Sun, S. , C. B. Murray, D. Weller, L. Folks, A. Moser. Science 287: 1989
(2000)
Tejada, J., X. X. Zhang and E. M. Chudnovsky, Phys. Rev. B 47: 14,977
(1993a)
( 1993a)
Tejada, J.,
J. , X.X.
X. X. Zhang and LL. Balcells. J. Appl. Phys. 73: 6709 (1993b)
Tejada, J. ,X. X. Zhang, L1. Balcells, O. Iglesias and B. Barbara. Europhys.
Lett., 22: 211(1993c)
Tejada, J., R. F. Ziolo and X. X. Zhang. Chemistry of Materials 8: 1784
( 1996); and references therein
Tejada, J., X. X. Zhang. E. del Barco. J. M. Hernandez and E. M.
Chudnovsky. Phys. Rev. Lett. 79: 1754 ( 1997a)
1754(1997a)
Tejada, J. , X. X. Zhang. J. M Hernandez, J. R. Friedman, M. P. Sarachik
and R. F. Ziolo. Magnetic Hysteresis in Novel Magnetic Materials Kluwer
Academic Publishers, Netherlands, p. 233( 1997b)
Thomas, L. , F. Lionti, R. Ballou, D. Gatteschi, R. Sessol
Sessolii and B. Barbara.
Nature 383: 145 ( 1996)
Vincent, E. J., Hamman, J. Prene and E. Tronc. J. Phys. France 4: 273
(1994)
Wang, Z. L. L.,, Z. R. DaiandS.H.
Dai and S. H. Sun. Adv. Mater. 12:
Adv.Mater. 1,944(2000)
12:1,944(2000)
Woods, S. I., J. R. Kirtley, Shouheng Sun and R. H. Koch. Phys. Rev.
Lett. 87: 137,205(2001)
Xiao, G. and C. L. Chien. Appl. Phys. Lett. 51: 1,280 (1987)
Yeshurun, Y., A. P. Malozemoff and A. Shaulov. Rev. Mod. Phys. 68: 911
(1996)
Zhang, Jinlong, C. Boyd and Weili Luo. Phys. Rev. Lett. 77: 390( 1996)
Zhang, X. X., L1. Balcells, J. M. Ruiz, J. L. Tholence, B. Barbara and J.
Tejada, J. Phys.: Conden. Matter 4: L163(1992)
L 163( 1992)
Zhang, X. X., J. Tejada, A. Roig, O. Nikolov and E. Molins. J. Magn.
Magn. Matt. 137: L235L235(( 1994)
Zhang, X. X.X.,, R. F. Ziolo, E.C.
E. C. Kroll, X. Bohigas and J. Tejada. J. Magn.
Magn. Matt. 140-145: 1,853(1995)
Zhang, X. X. , J. M. Hernandez, J. Tejada and R. F. Ziolo. Phys. Rev. B
54: 4, 10 l( 1996)
Magnetic Relaxation and Quantum Tunneling of Magnetization 181

Zhang, X. X. , J. M. Hernandez, E. del Barco, J. Tejada, A. Roig and E.


Molins,K. Wieghardt, J. Appl. Phys. 85: 5,633(1999)
Zhang, X. X. , In S. H. Yang and P. Sheng. eds. Physics and Chemistry of
Nano-structureed Materails. Taylor and Francis ltd, UK, p. 176 (2000)
Zhang, X. X. , H. L. Wei, Z. Q. Zhang and Lingyun Zhang. Phys. Rev. Lett.
87: 157,203(2001)
Zhang, X. X. , G. H. Wen, G Xiao and S. H. Sun, unpublished (2002)
Y. , M. P. Sarachik, J. Yoo and D. N. HendrickSon. Phys. Rev. B
Zhong, Y.,
62: R9256(2000)
Ziolo, R. F., E. P. Giannelis, B. A. Weinstein, M. P. 0' Horo, B. N.
Ganguly, V. Mehrotra, M. W. Russell and D. R. Huffman. Science 257:
219 (1992)

This work is supported by Hong Kong RGC grant HKUST6111 /98P. The author wishes to
thank Dr. H. L. Wei for the help in numerical simulation.
6 Nanostructured Exchange-Coupled Magnets

W. Liu, Y. Liu, R. Skomski and D. J. Sellmyer

6. 1 Introduction

In recent years, exchange-coupled magnetic nanostructures have attracted


much attention in the areas of permanent magnetism, magnetic recording,
sensors, soft magnetism and spin-electronics. The reason is that suitable
nanostructuring may improve the performance of an artificial material beyond
that of naturally occurring substances, real izing what is known as the
materials-by-design concept. This chapter focuses on permanent magnets,
whose ability to store magnetostatic energy is described by the energy
product. In the 20th century, energy product doubled every twelve years, and
the present-day record-holder Nd2 Fe14 B has energy products in excess of
451 kJ/
kJ/mm3. However, the outlook for discovering new ternary phases with
magnetizations significantly higher than Nd2 Fe14 B has been poor, and new
approaches are necessary if the energy product is ever to double again. For
example, as predicted in Skomski and Coey( Coey ( 1993), adding a soft material
with a high polarization, such as Fe65 C0 35 (JJo = 2.43 T), to an oriented
(lJo M ss =
hard magnet improves the energy product if the grain size of the soft regions is
sufficiently small. Compared to the present-day theoretical limit of 516 kJ/m 3
for single-phase Nd2Fe14 B, the energy product in suitably nanostructured Sm2
Fell N3/Fe65 C0 35 composites was predicted to be as high as 1090 kJ/ m3
Fe17
(Skomski and Coey, 1993). Related structures have been investigated by a
number of authors (AI-Omari and Sellmyer, 1995; Liu et al. , 1998a; Sawitchi
et al. , 2000; Bowden et al.,al. , 2000, Skomski et aI., 1999; Hadjipanayis,
1999; Fullerton et al. 1999)
1999)..
The development of exchange-coupled magnets has several starting
points. From a technological point of view, Coehoorn et al. (1988) first
exploited the remanence enhancement in isotropic Nd-Fe-B magnets; this
research has its scientific root in earlier random-field (Imry and Ma, 1975) and
random-anisotropy theories (Alben et al. , 1978; Chudnovsky et al. , 1986). A
second starting point is the investigation of magnetic multi layers (Nieber and
Kronmliller,
Kronmi..iller, 1989; Kneller and Hawig, 1991; Skomski and Coey, 1993),
which is now widely associated with Kneller's concept of exchange-spring
magnetism. Third, attempts to predict the nucleation-field coercivities He =
Nanostructured Exchange.Coupled
Exchange-Coupled Magnets 183

H N (H N is the nucleation field) for three-dimensional two-phase nanostructures


have given rise to a quantitative analysis of the permanent-magnet
performance of oriented two-phase nanostructures (Skomski, 1992; Skomski
and Coey, 1993). A nanostructure where this principle has been realized in
practice is FePt: the energy product of the composite exceeds that of hard-
magnetic FePt, in spite of the soft magnetic character of the magnetic phase
with a composition close to Fe3 Pt. In fact, the energy product of the
composite, 420 kJ/kJI m3 , approaches that of record-holding Nd22 Fel4 B and is
superior to that of FePt (Liu et ai.,
al. , 1998a).
As the best available permanent magnets, rare-earth-transition-metal
intermetallics have a lower saturation magnetization M s than many soft
magnetic materials. Furthermore, these materials are chemically very
reactive and also expensive due to a substantial rare-earth content.
Therefore, exchange-coupled composite permanent magnets are preferred,
consisting of both magnetically soft and hard phases, in which the former
provides a high saturation polarization and the latter contributes to a high
coercive field. Kneller and Hawig (1991) theoretically predicted a high
maximum magnetic energy product and an unusually high remanence ratio or
reduced remanence M,/ Mrl M s ' due to exchange coupling between nano-grains of
the soft and hard magnetic phases. In addition they predicted a unique
magnetic behavior characterized by a reversible demagnetization curve; that
is, a maximum recoil permeability as distinguished from the conventional
single-phase permanent magnets, where the demagnetization curves reflect
essentially the distribution of the irreversible switching fields. It is for this
unique magnetic behavior of the nanocomposites, in a sense resembling a
mechanical spring, that such magnets have been termed exchange-spring
magnets by Kneller and Hawig.
In nanomagnetism it is important to distinguish between isotropic and
aligned magnets. Isotropic structures are comparatively easy to produce, but
the remanent magnetization M, M r of randomly oriented grains with uniaxial
anisotropy is only half the saturation magnetization Ms. Since energy product
scales as M~, this amounts to an energy-product reduction by a factor of 4.
Intergranular exchange in isotropic magnets, realized by grain-size reduction,
improves the remanence by favoring parallel spin alignment in neighboring
MUlier et al. , 1990.
grains (Coehoorn et al. , 1988; Ding et al. , 1993a; Muller 1991). This
important feature is known as remanence enhancement. On the other hand,
intergranular exchange reduces the coercivity by averaging the anisotropy of
grains with different crystallite orientations. This random-anisotropy effect
(Imry and Ma, 1975; Callen et al. , 1977; Alben et al. , 1978; Chudnovsky et
al. , 1986) may overcompensate the benefit of the remanence enhancement,
and structures with very small grain sizes are actually used as soft magnets.
Typical examples are nanocrystalline Nd2 Fe14 B/Fe3B-Fe and Sm2 Fell Fel? N3/Fe
composites produced by melt-spinning (Coehoorn et al., aI., 1988) and
mechanical alloying (Ding et al. , 1993a), respectively. In oriented or aligned
184 W. Liu et al.

magnets, produced by various bulk and thin-film techniques, the crystallites


crystall ites
exhibit a complete or partial c-axis alignment. As a consequence, the
remanence is larger than MMs /2 even in the absence of exchange coupling, and
volume-averaged anisotropy remains nonzero. The benefit of nanostructuring is
the exploitation of the magnetization of a soft phase with high spontaneous
magnetization.
This chapter summarizes some key theoretical and experimental aspects
of exchange-coupled permanent magnets. Section 6.2 is an introduction to the
theory of exchange-coupled two-phase nanostructures, whereas Section 6. 3 is
devoted to experimental nanostructures including Fe-Pt (Section 6.3. 1), rare-
earth cobalt (Section 6. 3. 2) and Nd-Fe-B (Section 6. 3. 3). Finally, in
Section 6.4 we present some concluding remarks. Throughout this chapter we
will use the notation A : B to denote a nanocomposite material and AlB to
denote a multilayered material.

6. 2 Theory of Exchange-Coupled Magnets

An important theoretical problem is to derive magnetization curves by


simulating (Schrefl et al. ,1994) or modeling (Skomski aM and Coey, 1993) the
magnet's nanostructure. Although much progress has been made in the field of
numerical simulation (Sabiryanov and Jaswal, 1998a, 1998b; Schrefl et al. ,
1994; Schrefl and Fidler, 1998; Fischer and Kronmliller,
Kronmuller, 1998), some
problems are more conveniently tackled by analytic calculations. On the one
hand, the size of the simulated structures is limited to about 100 x 100 x 100
lattice points. In practice, this excludes phenomena such as layer-resolved
interatomic exchange on length scales smaller than 1 nm and the formation of
equilibrium domains on length scales larger than 100 nm. On the other hand, it
is often difficult to map large amounts of numerical data onto transparent
physical ideas.

6. 2. 1
6.2.1 Energy Product

A key figure of merit of permanent magnets is the energy product (BH )max'
which is twice the magnetostatic energy per magnet unit volume stored outside
the magnet. The energy product of hard-magnetic steel is about 1 kJ/m kJI m3 ,
whereas advanced Nd-Fe-B magnets have energy products of slightly more
than 400 kJ/m 3 . This means that less than 3 g of Nd-Fe-B are now able to
l-kg horseshoe magnet-a feature of major importance for advanced
replace a 1-kg
consumer electronics, car design, and computer technology.
In practice, the energy product is determ ined from the B-H hysteresis
determined
Nanostructured Exchange-Coupled Magnets 185

loop, which is obtained by plotting B = = /Jo


/-10 (H + M) (B is magnetic flux
density) as a function of H. In the B-H loop, the energy product is equal to the
maximum rectangular area fitting under the second quadrant of the loop
(Fig. 6. 1). The equivalence of the two definitions is shown by analyzing the
6.1).
integral fB. HdV, where the integration extends over the whole space. From
the magnetostatic equations V
V B = 0 and H = - V M
cf>M , where M
cf>M is the

magnetostatic potential, it follows that fB


B HdV =- f V
Hd V =- f(cf>MB)dV.
V (<f>MB)d f
V. This
integral is easily transformed into a surface integral over M
cf>M B, and since the
respective asymptotic radial dependences of M cf>M and B, 1/ R 2 and 1/ R 3 ,

4rrR 2 ,
overcompensate the effect of the surface area 41TR fB
B Hd V = O. Now we
= 0 outside the magnet and obtain the sought-for
take into account that M =

relation - LSide
LSide B Hd V f
= LU,Side
outside
2
v,
/-10 H d V, where the subscripts specify the
/Jo
volume of integration. Note that the energy depends on the demagnetizing field
and, therefore, on the shape of the magnet. In the case of low-coercivity
magnets, the energy product is maximized by cumbersome elongated or
horseshoe shapes, whereas advanced rare-earth transition-metal permanent
shapes are very compact.

h
~
'-' I /
~ /~
h
~
0 ------
J1J1oM(T)
oM(T) /
-.1----/ ------
: /
~ / / :b
:~
-1
-I

-I
-1 -0.5 0 0.5
J1
J1oH(T)
oH(T)
Figure 6. 1 Energy product (schematic). The energy product is equal to the gray area.

Energy product increases with coercivity He and remanence magnetization


M, but can never exceed the value /-IoM~/4 /JoM~/4 :s;;; /JoM~/4. However, if
~ /-IoM~/4.
magnetization were the only consideration, then aa iron with J s = /JoM /-IoM s =
2. 15 T would be used for permanent magnets with energy products as high as
920 kJ/m 3 . In fact, the coercivity of bcc iron is so low that energy products of
iron magnets are only of the order of 1 kJ/m 3 . The idea behind the
nanostructuring considered in this chapter is to maximize the magnetization by
adding a soft-magnetic high-magnetization phase without destroying the
coercivity.
186 W. Liu et al.

6.2.2
6. 2. 2 Fundamental Equations
The determination of the energy product requires the knowledge of the
hysteresis loop as a function of the magnets composition and microstructure.
As outlined in Chapter 1, the hysteretic behavior of a magnet derives from the
free-energy functional

F= ff{A[V(~)J
F {A [ V (~) -K,Kr- 1 (n~7)2
(n -
~7)2 -J1oM'H-~OM'Hd(M)}dV
PaM H - ~a M Hd(M) }dV
(6.1)

where M ss (r) is the spontaneous polarization, K, K 1 (r) is the first uniaxial


A (r) denotes the exchange stiffness, and nCr) is the unit
anisotropy constant, A(r)
vector of the local anisotropy direction. H is the external magnetic field, and
H d is the magnetostatic self-interaction field. For an arbitrary nanostructure,
the parameters entering Eq. (6. 1) are all local quantities. For example, the
anisotropy K,K 1 (r, T) is easily changed by varying the chemical composition
(Skomski and Coey, 1999b). In Eq. (6. 1) we have ignored higher-order
anisotropies, which are of secondary importance in the present context.
In isotropic magnets, nCr) is a random quantity obeying (n(r =0. This
gives rise to some specific random-anisotropy effects (Alben et al., 1978;
Chudnovsky et al., 1986; Sellmyer and Nafis, 1986), which are exploited,
for example, in Nd-Fe-B magnets (Coehoorn et al.. al., 1988) and briefly
discussed in Chapter 1. However, as analyzed in Skomsk i ( 1996), these
effects govern the ground-state behavior of random-anisotropy magnets but are
much less relevant to the phenomenon of hysteresis. In practice, the effect of
the random anisotropy can be studied by considering partly oriented (textured)
magnets, where n contains a random contribution but is close to the crystalline
J (r) remains close to J ss e z
c-axis (e z )' We also assume that the polarization J(r)
Us is saturaction polarization) until the reversal starts in the vicinity of He
(Skomski and Coey, 1993; Skomski and Coey, 1999b). This is reasonable
for nucleation-controlled reversal but underestimates M the c~ercivity
c~ercivity in M
pinning-controlled magnets (Section 6. 3. 3). As a consequence, we can
+ m)"by"=M
replace"= M s (J1-m 2 e zm)" by "=M ss (1-m 2 /2) e z + Msm"
M s m" and
"n=JT=CiT e z + a" by "n=(1-a
"n=.jT=(i2 "n= (1- 0 /2) e z + a." In the respective
2

equations, m and a are small transverse magnetization and easy-axis vector


components, both perpendicular to e z. z Figure 6. 2 illustrates the meaning of
these two quantities. Note that m=O means ideal spin alignment,M= M ss e z ,
and 0a = 0 means ideal c-axis ignment, n = ezz .
alignment,
c -ax is al
Putting the quadratic expressions for n as well as H = He z into Eq. (6. 1)
yields, up to a physically uninteresting shift of the zero-point energy,
Nanostructured Exchange-Coupled Magnets 187

f[
A(V m)2 + K 1 (m - a)2 -
= f[A(Vm)2+K1(m-a)2-
F = ~~JJoMsHm2JdV.
J.l oM s Hm 2 JdV. (6.2)

To ensure consistency with respect to the above approximations for M and n,


this equation contains only linear and quadratic terms. Furthermore, we have
assumed that the magnetostatic self-interaction can be incorporated into K 1
This is not always possible, but for the small length scales considered here the
corresponding corrections are of secondary importance.

Figure 6.2 Unit vectors nCr) and J(r)/


}Cr)/ JJ,s describing the polycrystalline easy axes and
magnetization. respectively.
the local magnetization,

From Eq. (6.2), the nucleation field is obtained by stability analysis.

Writing the energy functional as F = = f ryd V, equilibrium is given by the


1Jd

= O. Explicitly (Skomski and Coey, 1999),


functional derivative 5F /5m(r) =

~--V(
~ =_ V (
5m(r) - dV
dry
d1J
V mer)
)+-.i!L
)+~
dm(r)
(6.3)

and (Skomski et aI.,


al. , 1998a)

- V [A (r) V m]
[A(r) + [K 11 (r) - J.l
JJ o0 M ssH
H/2]m = K 11
/2] m = K a. (6.4)

The spatial variation of A (r) is important for some grain-boundary problems


~
(Section 6. 3). For the usually considered two-phase magnets, A (r) ::=::::
10 pJ/m throughout the magnet, the effect VA
V A (r) can be ignored for
V [A(r)Vm]=AV 22m.
qualitative analysis, and V[A(r)Vm]=AV
In Eq. (6. 4), the x and y components of m = mx ex
ex = + my ey
ey are
decoupled and degenerate, so that the vector m (r) can be replaced by any
unspecified nucleation mode If (r). It is, however, convenient to think of
'f(r).
If(r) as the magnetization m x (r) in x direction. Note that the random
'fer)
contribution to the easy-axis direction, o(r),
a (r), acts as a perturbation tending to
misalign the spins but does not affect the eigenmodes of Eq. (6.4). In other
words, in lowest order the nucleation field depends on the strength of the
anisotropy, K 1 (r),
(r). but not on the easy-axis misalignment.
188 W. Liuetal.
Liu et al.

6.2.3 Nucleation Field


= 0) and making the fair assumption
Considering perfectly aligned magnets (a =
= const. throughout the magnet, we can rewrite Eq. (6.4) as
that A =
V2m
A '1
- A m + - IJ o M s H/2) m
[K 1 (r) -lJ = O. (6.5)

where the use of m = = I m I is justified due to the resulting degeneracy with


1 1

respect to the direction of m. Equation (6. 5) has the same structure as the
single-particle Schrodinger equation

- ~~ V 2 'f'+ [V,(r) - E]'f' = O. (6.6)

This quantum-mechanical analogy makes it possible to use ideas known from


quantum mechanics to solve micromagnetic problems. In particular, the
nucleation field and the nucleation mode are analogous to the ground-state
energy and the ground-state wave function, respectively.
To illustrate this analogy, we consider a soft-magnetic cube of volume L6
L5
embedded in a very hard matrix (Fig. 6.3).
6. 3). The corresponding wave functions
are the particle-in-a-box states for an infinite potential well

= 'f'o
'f'(x,y,z) = sin(1T~xX)sin(1T~YY)sin(1T~zZ)
sin(n~xX)sin(n~YY)sin(n~zZ) (6.7)

31T 2 n
and the ground-state energy is 3n h. 2 /2mL 2. Using this energy and comparing
(6. 5) and (6.6)
Eqs. (6.5) (6. 6) we obtain the nucleation field

(6.8)

where M s is the spontaneous magnetization of the soft phase. This result can
also be written as

H
N
= H 35~Mh (6.9)
L 2 Ms

where H.H a = 2K,/IJ


2K 1 /lJ o M s is the anisotropy field of the hard phase, M h is the
magnetization of the hard phase, and 5 = 1T(
OsB = n( A / K 1, )) '/2
1/2 is the Bloch-wall width

of the hard phase.


From Eqs. (6.8) and (6.9) we see that H N increases with decreasing
o. However, in reality it never goes to infinity, because the
inclusion size L o
anisotropy field H.H a of the hard phase provides a cut-off, which is ignored in
(6.7)-
Eqs. (6.7) - (6.9). However, extrapolating Eq. (6.9) (6. 9) to H HN~ Ha reveals
N """" H.

that the nucleation field reaches the anisotropy field of the hard phase when the
size or radius of the soft inclusion becomes comparable to the Bloch-wall width
of the hard phase. This rule has been specified for a variety of geometries and
parameter sets (Skomski, 1992; Skomski and Coey, 1993). Since the Bloch-
Nanostructured Exchange-Coupled Magnets 189

Figure 6. 3 Soft-magnetic cube embedded in a very hard single-crystalline matrix (K 1


M; ).
Po M;).
IJo

wall width of typical permanent magnets is of the order of 5 nm, soft regions
should be smaller than about 10 nm to ensure complete exchange coupling.
However, this value depends to some extent on the geometry of the
nanostructure; it is somewhat larger for multi layers and somewhat smaller for
ideal spheres. Furthermore, there are magnetostatic self-energy contributions
which cannot be incorporated into K l ' These contributions give rise to
corrections whose size depends on the orientation of aspherical soft regions.
The fact that the switching field of the soft phase is comparable to the
required coercivity, rather than comparable to the anisotropy of the hard
phase, leads to somewhat larger critical sizes (Skomski et al. , 1999). On the
other hand, nanostructuring cannot be used to significantly improve the Curie
temperature of a constituent phase (Skomski and Sellmyer, 2000). In contrast
to the domain-wall width of the hard phase, that of the domain-wall width of
the soft phase is irrelevant to the problem. Soft-magnetic materials are
characterized by wall widths exceeding 100 nm or even 1 IJm, but this length is
of no consequence in the context of hard-soft permanent magnets. As a matter
= 0 in the soft-
of fact, in almost all cases of interest it is safe to put K 1 =
magnetic phase, corresponding to an infinite wall width. Some other aspects
of this exchange-length problem are discussed in Chapter 1.

6.2.4 Energy Product


When the sizes or radii of the soft phase are smaller than os,
OB' then H N as a
function of L o0 reaches a plateau. The quantum-mechanical equivalent is the
lowest-order perturbation theory (virtual-crystal approximation), where the
potential energy V ( r) is averaged over atomic disorder. In this plateau
regime, both the anisotropy and the magnetization of the composites are equal
to the corresponding volume averages, and the maximum energy product
(Skomski and Coey, 1993) is
190 W. Liuetal.

( BH) max =-l4 110 M2(1_110(Ms-Mh)Ms)


s 2K
(6.10)
h

This energy-product value is achieved when the volume fraction of the hard
phase is equal to M; /4110 K h' Since (anisotropy constant of hard magnetic
phase) K h is very large in advanced permanent magnets, the second term in
the bracket is small so the energy product approaches the ultimate value of
110 M; /4. As analyzed by Skomski and Coey (1993), for suitable combinations
of hard and soft phases, Eq. (6. 10) predicts energy products of up to about
1000 kJ/
kJ/mm3 . Note that the nonlinear
nonl inear dependence of the energy product on the
magnetization is an important aspect of two-phase permanent magnets. In the
case of a linear dependence, the maximum energy product of a hard-soft
nanocomposite would be intermediate between the energy products of the hard
and soft phases.
The val idity of Eq. (6. 10) is independent of the shape of the soft and hard
validity
regions so long as the size of the soft regions remains in the plateau limit. The
M-H hysteresis loops corresponding to Eq. (6. 10) are ideal rectangular loops.
Large soft regions lead to inflections (shoulders) in the hysteresis loops (Liu et
aJ. , 1999a) and to a significant reduction in energy product and coercivity.
al.
Furthermore, structural imperfections in real magnets, such as imperfect
crystallite alignment, give rise to a smoothing of the loop and further reduce
the energy product. It is therefore of key importance to control the
nanostructure of the composites.

6.2.5 Micromagnetic Localization

A specific feature of nucleation modes in inhomogeneous magnets is their


(Skomsk i, 1998). By contrast, the well-known coherent-rotation
local ization (Skomski,
curling
and curl ing modes are delocalized,
delocal ized, that is, they extend throughout the
magnet. Somewhat simplifying, polarization reversal tends to start at
locations where the" local" anisotropy field H
the "local" H a (r) = = 2K 11 (r) /110o M s (r) is
(r)/l1
lowest, but exchange coupl ing to highly anisotropic regions which means that
intrinsic coercivity M He> min [H a (r)]. Figure 6. 4 illustrates the
delocalization mechanism for a double well: due to the finite anisotropy of the
hard phase, the nucleation mode penetrates into the hard phase. The quantum-
mechanical analog of this penetration is tunneling.
tunnel ing. As quantum-mechanical
tunneling leads to hybridization and to a reduction of the ground-state energy,
micromagnetic delocalization reduces the nucleation field. This is the reason
local ized model Eqs. (6.
why the simple localized ( 6. 7) - (6. 9) overestimates the
nucleation field for small cube sizes.
The micromagnetic (or "nanomagnetic") localizationlocal ization problem is very
similar to the problem of electron localization in disordered solids. The degree
of localization depends on the sizes of the soft and hard regions, on the
anisotropy of the hard phase, and on the dimensionality of the problem. In one
Nanostructured Exchange-Coupled Magnets 191

Hard Soft Hard Soft Hard


x
(a)

(b)

6.44 Delocalized nucleation mode for two parallel soft layers in a hard matrix:
Figure 6.
(a) anisotropy profile and (b) nucleation mode m.

dimension, arbitrarily small disorder causes the localization of all eigenmodes


(Anderson, 1958). In the magnetic analogy, this case has been analyzed for
long (1 IJm) and thin (less than 10 nm) transition-metal nanowires (Skomski
et aI.,
al. , 1999; Sellmyer et al. , 2001).

6. 2. 6 Texture
The perfect c-axis alignment is difficult to realize in practice, and most
nanostructures, such those investigated in (Liu et aI., 1998a), exhibit a
nonnegligible degree of crystalline texture. The starting point for the
description of weak texture is Eq. (6.4),
(6. 4), where the polycrystalline easy-axis
disorder a (z) acts as a random inhomogenity. For A (r) = = const. and soft
sUfficiently small to ensure a plateau behavior on a local scale,
regions sufficiently
Eq. (6.4) yields (Liu et al. , 1999a)

m2 > =
( m2)
< 4K~v
4 K ~v < 2>
2(a2) 8K~vA
8 K ~v A (
< VV 22 >) (6.11)
(2K av + J.1/JoM H) a
o M s H)2 (2K av + J.1
/Joo M H)3 a'
MssH)3 a
where average anisotropy K av = = fhK h . Since, in lowest order, M == M s (1-
<m2 >) /2), this equation can be used to discuss hysteresis loops. Figure 6.5
(m
shows that the result for two geometries. A consequence of these equations,
which go beyond the ground-state properties considered, e. g. by Chudnovsky
et al. (1986), is a secondary reduction of the energy product proportional to
<a2 >) (Skomski et al. ,1998).
(a
The texture of a nanocrystal Iine
line magnet is closely related to the
remanence enhancement and other random-anisotropy effects. Randomly
oriented grains exhibit a remanence that depends on the anisotropy of the
192 Liu et al.
W. Liuetal.

d=3(bulk)

d=2(thin film) 1::


/ 1
I ~
~
,::J
I
I I
I I
II
-6 -51
-5 -4 -3 -2 -I 0
II llo H(T)
poH(T)
I
Figure 6. 5 Strong-coupling hysteresis loops for two and three-dimensional magnets.

grains. For example, iron-type (K 1 > 0) and nickel-type (K 1 < 0) cubic


magnets, exhibit Mr/M s values of O. 832 and O. 866, respectively, as
compared to Mr = M Mj2
s/2 for uniaxial anisotropy. Exchange coupling enhances
the remanence beyond these values. In terms of Eq. (6. 11), we exploit that
a V 2a = V (a Va) - (Va)(Va)22 and find, after short calculation, that the
remanence increases with increasing exchange stiffness A. The small
parameter of this theory is A / K av L 2, where L is the average grain size.

6.2.7
6. 2. 7 Effective Exchange

Exchange effects at grain boundaries are of importance in permanent


magnetism, magnetotransport and magnetic recording. Pronounced
intergranular exchange is undesirable in magnetic recording, because it leads
to cooperative "interaction domains" and reduces the storage density
(Fukunaga and Inoue, 1992; Sellmyer et al. , 1998). On the other hand, two-
phase permanent magnetism relies on a strong exchange coupling between
hard and soft regions (Skomski and Coey, 1993; Skomski and Coey, 1999b).
The exchange energy between two adjacent grains II and 1I is n
proportional to Jeff S I S n
H , where S I and sHare
snare the magnetization directions
in the centers of the grains, and Jeff is an effective exchange constant. A
popular approach is to assume that Jeff = = NJ o , where N is the number of pairs
of adjacent atoms and J o cc k 8B T c is the interatomic exchange constant.
However, this approximation greatly overestimates the intergranular exchange
(Skomski,
( Skomski, 2001; Skomski et al., 1998), because it assumes that the
exchange energy is stored in one atomic layer. A qualitative scaling argument
is based on the exchange term fA (VM/M
(V'M/M s )2dV, which yields Jeff~ ALo~
1 2
N /J o , where L is the grain size. This estimate is more realistic than the
above-mentioned pair model and yields a significantly reduced effective
Nanostructured Exchange-Coupled Magnets 193

exchange, but it is unable to account for sharp grain boundaries and anisotropy
effects.
For an .ideal
ideal interface between semi-infinite hard and soft grain, the
exchange problem has been solved in (Skomski et aI., 1998a). Jeff Jeff is

obtained by considering the case SIS I :::::::: ~ e z'


~ Ssnn :::::::: z , so that Eq. (6.4)
(6. 4) can be
used. In the soft and hard regions the solutions of Eq. (6.4) are ml m, + m2 exp
s /A)1/2 x] and m3 + m4 exp[- (K h
[(KjA)1/2
[(K /A)l/2 x], respectively, and
h /A)1/2

integration over the corresponding energy density Eq. (6.2) (6. 2) yields (Skomski
et al.,
al. , 1998a)

(6. 12)

where So is the grain-boundary area, K hand K s are the anisotropies constant


of the hard and soft grains, respectively, and A is the common exchange
6.66 shows the corresponding magnetization
stiffness of the two grains. Figure 6.
profile m(x) and the energy density. From Eq. (6.12) we see that Jeff =0 for
K s = 0, corresponding to a free switching of the soft grain. This unphysical
limit is a consequence of the assumed semi-infinite character of the grains.
Analyzing the problem for finite-size soft grains reveals that Eq. (6. 12)
remains val
valid
id if one uses the"
the "effective"
effective" anisotropy K s = 4
4AA/ d
L~ .

"o<::
u.~ o
.0 Soft phase Hard phase
"["
.[ .;;;
Vi
<::
"o<:: "
.g
<l)
"0
.~
N
N
~
""
<l)

.~
.~ <::
Soft phase Hard phase U-l
W
~
~ oL-----------~x
o x

6. 6 Grain interface between soft (K s5 = 0.25


Figure 6.6 MJ/ m3 ) and hard (K h = 4 MJ/
O. 25 MJ/m MJ/mm3 ) :

(a) spin orientation and (b) energy distribution 1)


I] (x) in arbitrary units.

Up until now we have assumed that A (r) = const. throughout the


material. This is a fair approximation in many cases, because A is of the
order of 10 pJ/m for most ferromagnetic materials of interest. However,
chemical and structural disorder may yield a strong reduction of A. This affects
the intergranular exchange and leads, for A = 0, to Jeff Jeff = O. Figure 6.7 shows

two misaligned grains, characterized by nI ~ n


n I :::::::: ~ ez and exhibiting a
nnn ::::::::
reduced interatomic exchange (J') at the interface. We will investigate the
spin structure and coupling behavior of this structure by both continuum and
layer-resolved methods (Skomski, 2001; Skomski et al. , 2001).
On a continuum model, exchange inhomogeneities are described by the
"V (A "m)
V m) term in Eq. (6.4).
(6. 4). As discussed in (Skomski and Coey, 1993),
this equation amounts to the boundary condition.
194 W. Liu et al.

'~
x

Figure 6. 7 Two neighboring grains and grain boundary (Skomski et al. , 2001). The J
and J' are interatomic exchange constants between atoms in different layers, and the
thickness of the grain-boundary region is t "'" a. On a continuum level, J' / J = A' / A .

[A(X)amJI
ax X
o-'
= [A(X)amJI
ax X
o+'
. (6.13)

Figure 6.8 illustrates the physical meaning of this boundary condition. A jump
in A (x) yields a change in the slope of the perpendicular magnetization
component m (x). However, when the exchange stiffness in the grain-
boundary region is much lower than that of the two adjacent phases, one
encounters a quasi-discontinuity of the magnetization. Note that this phenomenon
does not depend on the hard or soft character of the adjacent phases.

Hard Soft

large A

x
(a) (b) (c)

Figure 6. 8 Boundary conditions and exchange: (a) hard-soft interface with common A ,
(b) interface between two ferromagnetic phases with different A, and (c) quasi-
discontinuous magnetization due to strongly reduced exchange in a grain-boundary region
between grains II and II .n.
Applying Eq. (6. 4) to the structure of Fig. 6. 7 yields, in analogy to
Eq. (6. 12), the effective exchange

Jeff = So J AK h 1 (6. 14)


1 + lftA
lTtA
2osA'
= K hh in Eq. (6.12)
By putting K ss = (6. 12) we see that the expression So (AK
( AK hh))1/2
1/2
Nanostructured Exchange.Coupled Magnets 195

describes an ideal interface. Analyzing the denominator in Eq. (6.14) reveals


small values of the interface exchange J' lead to small values of Jeff' J eft , but since
~ a 50'
t ::::::::: 00' the effect of J' is disproportiontely small. Note that electronic
structure calculations on FePt/Fe interfaces (Sabiryanov and Jaswal, Jaswal. 1998a)
~ 50 %, corresponding to a reduction of Jeff
predict J' / J ::::::::: J eft by only somewhat
more than 10%. Thus, moderate reductions of the interatomic exchange J' do
not lead to a decoupling of the grains.
The relative strength of the quasi-discontinuity is

(6. 15)
.1 =
Ll = 1 + 2A'os
2A'OB .
TIAt
nAt

For zero exchange in the grain-boundary region, A' = = 0, this equation yields
an ideal magnetization jump (Ll(.1 = 1), whereas for t = 0 the quasi-discontinuity
vanishes (Ll(.1 = 0). As in the case of the effective exchange, the small ratio
OB tends to suppress the influence of the reduced interface exchange.
t / Os
In a layer-resolved analysis, Eq. (6.4) must be replaced by

n, n+ 1 ( m n
N J n. - m n+ 1) + K 1 Sot 0 m n = K 1 Sot 0 a n (6.16)

n + 1 ~ A (r) to is the interlayer exchange coupling


where IJ nn,,n+l::::::::: coupl ing between adjacent
atoms in the n-th and (n + 1)-th layers, and each layer (index n) has a
thickness to = a. The main assumption is that the interatomic interface
exchange J O,l = J' is smaller than the bulk exchange J n,n+ 1 = J (J is exchange
constant). A typical solution of Eq. (6.13) is shown in Fig.6.9. Compared to
the continuum theory, there are only minor corrections. For example, as in the
continuum case, the magnetization perturbation decays exponentially in the
grains, but the corresponding decay length

mr

.
~
l:: \ ]
mil
~
-10 1 2
n

6. 9 Spin structure in the vicinity of the grain boundary. The jump !:lm
Figure 6.9 I:lm means a
quasi-discontinuity of the magnetization at the grain boundary showing the normalized
magnetization projection m n as a function of the layer index (layer number n).
196 W. Liu et al.

II = 1 C6.17)
arcos h Cl + K,S o t o/2J)

is slightly different from the continuum expression}.


expression II = CJ / K 1 So t) 1/2.

6.3
6. 3 Experimental Systems

Experimental and theoretical research on nanostructured exchange-coupled


magnets has been active since 1988. Here, we will review the structure and
magnetic properties of the nanocomposite magnets including some new results
and phenomena, from an experimental point of view. Based on the hard phase
in the nanocomposite magnets, three types of the magnets will be reviewed:
CDFePt-based magnets, ~rare-earth-cobalt-based
CVrare-earth-cobalt-based magnets, and @ Nd-Fe-B-
based magnets. For each of them, the phase diagrams, crystal structure,
intrinsic properties related to the hard magnetic compound as a component of
the nanocomposite and the structure and magnetic properties of the
corresponding permanent magnets are briefly reviewed, and then some
properties of the nanocomposite or nanostructured exchange-coupled magnets
are described in detail. Both bulk and thin-film nanocomposite magnets will be
described. Depending on processing methods, permanent magnets can be
divided into isotropic and anisotropic ones. Most of nanostructured bulk
magnets are isotropic, including nominally single-phase and composite
magnets, but for many of the thin-film magnets the hard magnetic grains are
oriented to a certain extent.

6.3.1 FePt-Based Magnets


6. 3. 1. 1
6.3. Phase Diagrams and Crystal Structures
The equilibrium phase diagram of Fe-Pt binary alloy system given by Massalski
C1990a) is shown in Fig.6. 10. It contains the following regions: CD liquid, L;
~ bcc terminal solid solution Co-Fe); @ face-centered cubic (fcc)
CV Cfcc) continuous
solid solution, y-CFe, Pt) or y;
y-CFe,Pt) CD bcc terminal solid solution, Cex-Fe);
@ cubic AUCU3 -type Fe3 Pt; @ tetragonal AuCu-type FePt; (j) cubic AuCur
AuCurtype AUCU3-
type FePt 3 ; bct martensite phase, ex'; and fct martensite phase, y'.
The Fe-Pt binary alloy system has a complete solid solution range in a fairly
wide temperature range. When quenched from high temperatures, the crystal
structure of these alloys is a disordered fcc structure. However, when
annealed at lower temperatures, both stoichiometric and nonstoichiometric
alloys with various degrees of order can be prepared. Fe-Pt crystal structure
datais
data is concluded in Table 6. 1.
Nanostructured Exchange-Coupled Magnets 197

Weight Percent Platinum


010203040 50 60 70 80 90 100
1800t-'-~-~"""""'--+---+--~~~-~~~-~-----I1769
1800~~-~"""""'--+---+----~~~-~~~-~----;1769 c
'C

L
1600 1538 c
2.5
1519C
---
(O-Fe)
14001394C
~
~ (Y- Fe,Pt)
(Y-Fe,Pt)
:I
::l
E!
' 1200
"'"0-
0.,
E
~
1000 912 C
912'C
Magnetic

70 80 90 100
Pt
Atomic percent platinum
Figure 6. 10 Phase diagram of the Fe-Pt system CMassalski, 1990a).

Table 6.1 Fe-Pt crystal structure data CMassalski, 1990a).


Composition, Pearson Space Strukturbericht
Phase Prototype
PtCat. %) symbol group designation

Co-Fe)
C6-Fe) o to 2.3 cl2 Im3m A2 W

Cy-Fe, PI)
PO o to 100 cF4 Fm3m A1
Al Cu

Cex-Fe)
Ca-Fe) o to ? cl2 Im3m A2 W

ex
a 22 to 24 cl2

Fe3Pt
Fe3 Pt 16 to 23 cP4 Pm3m L 12 AUCU3

Y 24 to 26 tP4

FePt 35 to 55 tP2 P4/mmm L1 o0


Ll AuCu (I)

FePh 57 to 79 cP4 Pm3m L112


L AUCU3
AUCU3

The phase diagram Fig. 6. 10 shows that both the AuCu3 type (L 12 ) and
the AuCu type (L 10 ) ordered phases exist. The ordered FePt3 is
antiferromagnetic while both the ordered FePt and PtFe3 are ferromagnetic. As
a famous Invar alloy, PtFe3 shows very strong anomalies in the thermal
expansion coefficient and spontaneous volume magnetostriction (Hayase et
al. , 1971). The structural and magnetic phase diagram of bulk Fel-
aI., Fe 1 - .x Pt.
Pt x in a
small concentration range (0. 22 < x < O. 32) is shown in Fig. 6. 11
(Rellinghaus et al. , 1995; Sumiyama et al. 1983). The fcc phase exists in the
198 W. Liu et at.
al.

Fe,- x Pt x alloys at temperatures above T = 400 K, whereas at low


Fe,- x Pt x alloys with concentrations x < O. 28 undergo a
temperatures, the Fel-
martensitic transformation from the fcc (y) into the bcc (oJ(0:) phase. For higher
concentrations, the Fel- x Pt x alloys retain the fcc structure to low
temperatures. The Fel-
Fe,- x Pt x alloys can exist in the ordered state as well as in
the disordered state.
700

600

500

o'---Z:C:
L----,2~2,------;:2'":-4
Z'---:::Z'7 --,Z:':6,.-----:::Z"'"g-3:C:0,---:::3'=-Z-
4 ---,2::':6,.------;:2"'"8-3::':0,-----::3"'"2-
Pt concentration x (at.%)

Figure 6. 11 Structure and magnetic phase-diagram for Fe,-


Fe, _ x Pt x alloys with O. 22 < x
< 0.32
O. 32 (Rellinghaus et al.
aI.,, 1995; Sumiyama et al.,
al. , 1983).

The L 10 phase shown in Fig. 6. 12 has a superlattice structure consisting of


mono-atomic Fe and Pt layers stacked alternately along the c axis, which is
the magnetic easy axis. The axial ratio cia c / a is slightly less than one, ccla~
/ a~
o.O. 96 - O. 98. In the disordered fcc phase, the Fe and Pt atoms randomly
occupy the lattice sites. The long-range Bragg-Williams order parameter for
the ordered phase is defined as (Warren, 1990; Cebollada et al. , 1994):
= (r Fe
S = - X
X Fe ) / I Y PI == (r PI - X PI )
X I/ Y Fe ( 6. 18)
where x Fe(Pt)
Fe<Ptl is the atomic fraction of Fe or Pt in the sample, Y Fe(Pt)
Fe<Ptl is the

fraction of Fe or Pt sites, and rrFe(Pt)


Fe<Ptl is the fraction of Fe or Pt sites occupied by

the correct atoms. S is equal to one in perfectly ordered films of the

Figure 6. 12 Crystal structure of CuAu (I -type ordered FePt alloy (Watanabe et al. ,
(I))-type
1996).
Nanostructured Exchange-Coupled Magnets 199

stoichiometric composition Feso


Fe50 Pt 50
so and is zero for a chemically disordered
film. The long-range order parameter can be determined from the ratio of the
integrated areas of the super
superlattice
lattice peaks over the fundamental peaks of the X-
ray diffraction pattern for random alloys (Rudman and Averbach., 1957;
Cebollada et al. , 1994). However, for thin films, the calculation of the long-
range order parameter from the XRD pattern is more complicated due to the
existence of certain texture. In practice, S8 can be estimated by (Watanabe et
al. , 1996)

As! A, = (A s/ A , )s=1 8 2 . (6.19)

Here As and A AIf are the integrated areas of the superlattice peak and the
fundamental peak, respectively; (As!(As/A, )5=1 is the value of As/A,
AI )s=1 As/AI for a
perfectly ordered sample (8 (S = = 1), which can be obtained by long-time
annealing of the disordered alloys. High magnetocrystalline anisotropy of the
compound FePt originates from the chemically modulated superlattice
super lattice of Fe
and Pt atoms. The anisotropy depends on the compositional deviation from the
equiatomic stoichiometry, on S 8 and on the tetragonal distortion (Sakuma,
1994) .
1994).

6.3.1.2 Magnetic Properties of Fe-Pt Alloys


The high-temperature phase with fcc structure is practically useless for
permanent-magnet applications because its magnetic anisotropy is low.
However, when rapidly cooled solid-solution Fel-. Fe1-. Pt. alloys are annealed at
lower temperature, the transformation of the disordered fcc phase to an
ordered face-centered tetragonal (fct) phase around the equiatomic
composition results in a high easy-axis anisotropy. For the ordered alloy FePt,
which crystallizes in the fct L 10 0 structure, the magnetocrystalline anisotropy
3
constant K 1 1 (> 1 MJ/m ) , and the magnetic anisotropy field lJo H a is more than
J.Jo Hais
10
lOT T (Ivanov et al. , 1973; Watanabe et al. , 1996). An anisotropy constant of
about 10 MJ/m3 has been found in the molecular beam epitaxy (MBE)-grown
FePt LlL1 o0 films (Farrow et al.,
aI., 1996a). FePt thin films have attracted much
attention because of their potential appl ication as recording media (Luo and
application
Sellmyer, 1995; Coffey et al., aI., 1995). The magnetic properties of the FePt
alloys have been investigated extensively. For example, a high coercivity of
0.43 T (Fig.6. 13) and a maximum energy product (BH) max max of 159.2 kJ/m 3
were reported in an off-stoichiometric Fe-Pt alloy containing 38.5 38. 5 at. % Pt
(Watanabe and Masumoto, 1983, 1985).
Stoichiometric FePt alloy in a bulk form does not show high magnetic
hardness, because the microstructure in bulk systems is very different from the
thin films. In addition, bulk FePt is quite expensive and therefore not used
industrially. Current interest in this alloy system is mainly concerned with basic
physics or applications of magnetic thin films. Since K 1 1 is much higher than the

shape anisotropy of thin films, controlling the crystallographic orientation of


thin films during crystal growth yields perpendicular anisotropy (Mcintyre et
200 W. Liu et al.

8 0 Thin film(After M.Watanabe


M. Watanabe et al.)
o Bulk (After K.Watanabe et al.)
7
~6
o'::"45
u
::t:: 3
2
I
~LO--"-=-_I..L.-~~---,L--,L.---'60
60

Figure 6. 13 Coercivity of FePt alloys and FePt


Fept (001) thin films as a function of alloy
composition. Data of bulk alloy magnets are after Watanabe and Masumoto (1983) and
those of thin films are after Watanabe and Homma (1996).

al.,
al. , 1995, 1997). In this respect, the magnetic properties of the FePt thin
films prepared by sputtering (Visokay and Sinclair, 1995; Watanabe et al. ,
1996) and MBE technique (Mitani et aI., al. , 1995; Farrow et ai.,
aI., 1996b) were
studied. Hong et al. (1998) reported that the optimum hard magnetic
properties are observed in the slightly Fe-rich side of Fe-47. 3 at. % Pt alloy
with a Pt buffer layer of 100 nm thickness, i. e., (BH)max = = 285 kJ/m 3 and
J..IoMH c = 0.69 T. These values are almost two times higher than the best
1J0MH
values reported in bulk magnets (Fig.6.13).
(F ig. 6. 13). Goto et al. (1999) reported that
an expitaxial FePt film prepared by electron beam evaporation with a granular
structure for the thickness of 120 A exhibits a huge coercivity of 5.4 T at room
temperature. Due to the high anisotropy, FePt is an interesting candidate for
permanent magnetic applications (Franklin et al. ,,1954;
1954; Craik et al.,
aI., 1972;
Tanaka et al. , 1997).

6.3.1.3 Nanostructure Synthesis


A high coercivity can be obtained in FePt films due to its high uniaxial
magnetocrystalline
magnetocrystall ine anisotropy, but the saturation magnetization of the FePt
L 10 phase is lower than that of the disordered fcc Fe-rich and bcc Fe-rich
phases. In order to increase the energy product of the FePt magnets, it is
possible to introduce a magnetically soft phase with high saturation
magnetization into the sample to enhance the remanent magnetization via the
intergrain exchange coupling (Skomski and Coey, 1993). For synthesizing
exchange-coupled permanent magnets, the following three factors should be
taken into account in the structure and composition of the magnets: CD a hard
phase with a sufficiently high uniaxial magnetocrystalline anisotropy for
developing high coercivity and a soft phase with a sufficiently high
magnetization for developing high energy product should be selected; (?) If
heat treatment is necessary, the hard phase and the soft phase should be able
to co-exist and with an inter-phase coherent lattice structure; @ for an
effective exchange coupling between the hard phase and the soft phase, the
Nanostructured Exchange-Coupled Magnets 201

size of the soft phase must not exceed about twice the domain wall thickness of
the hard phase. In the tetragonal FePt phase, the wall thickness is about
4 nm, which means that the grain size of the soft phase in the exchange-
coupled Fe-Pt system should be no greater than about 8 nm. Sabiryanov and
Jaswal (1998a) performed a first-principle calculation leading to an energy
product of 716.4 kJ/m 3 in a perfect FePt/Fe two-phase structure.
Xiao et al. (2001 a, 2001 b) investigated the effect of remanence
Fe59 75 Ph95 Nbo.75 bulk alloys annealed at 625C for different
enhancement of Fe5975
times. An almost single-phase disordered fcc phase can be obtained after
quenching in ice water from 1325 C. Two-phase nanocomposite bulk samples
can be obtained by means of annealing for different times (5 min - 110 h) at
625 C. It was found that the intrinsic coercivity M He continuously increased
with increasing annealing time up to 80 h. An excessively long time of
annealing resulted in decreasing of the coercivity, but the saturation
magnetization M s was relatively little affected by annealing time. For the Nb
doped Fe-Pt system, Xiao et al. explained the experimental results based on
two assumptions: CD the magnetocrystalline anisotropy of the hard magnetic
fct-phase particles increased with annealing time; (2) ~ there was a distribution
of the hard magnetic phase particles having different anisotropy values, at
least for the shorter annealing time. Thus, based on the assumption CD, with
increasing annealing time, the anisotropy constant (of hard phase) K k
increases, which results in a continuous decrease of the critical grain size
b em
em = (A m / K 1 ) 1/2 of the soft magnetic phase. An interesting result
1T (Ami
observed in the work of Xiao et al. is that the measured reversibility of
magnetization for various annealing time remains lower than theoretically
predicted by Kneller and Hawig (1991). This behavior finds its origin in the
distribution of regions with different anisotropy values, based on assumption
~. If M
(2). Mree and M tot denote the recoil and total magnetization, the exchange
Mlot
spring behavior with M Mreel
ree / M tot values close to unity is expected only when the
MIOI
demagnetizing field is smaller than the nucleation field H N at which domain
walls will penetrate into the hard magnetic phase and cause irreversible
losses. In the system studied by Xiao et al. , this condition may not be met,
and even small demagnetizing fields will lead to irreversible losses and cause
the total measured value of M ree I M tot to be low. The alloys annealed for
Mree/Mlol
shorter times are the most vulnerable in this respect because of the large
fraction of regions with low K 1I values. For longer annealing times, this fraction
Mrecl Mtot is shifted to higher values. Bruck et al. (2001)
becomes smaller, MreciMtot
studied the influence of the disorder-order transformation on remanence
enhancement and coercivity in Fe-Pt based materials by changing the
anneal ing temperature and time, and the type of additives. The soft and hard
phases are observed in the samples depending on the annealing process. Hard
magnetic properties of lJoM /Jo M s = 0.98T,
0.98 T, IJOM/JOM He = 0.37T
0.37 T and (BH ) max ==
(BH)max
125 kJ/m 3 in (Fe
(Feo606 Pt o
044)995 Nb
)995Nbo5 o 5 have been achieved.
Theoretically, an ideal case should be that the hard phase aligns with
magnetic field and the soft phase is embedded in the hard phase. In practice,
202 W. Liu et al.

such an ideal nanostructure is very difficult to realize. Several processing


methods have been employed to synthesize exchange-coupled magnets.
Physical vapor-deposition methods such as gas evaporation, laser ablation,
sputtering, and molecular-beam epitaxy all can be used to synthesize
nanostructured materials. Other methods such as mechanical alloying, rapid
quenching, etc. also can be used to generate nanostructured materials.
Among these methods, sputtering is cost effective, has the advantages of
controllable deposition rate and therefore has been successfully used to
synthesize various nanocomposite magnets.
Two approaches were attempted to realize the nanocomposite of
FePt:
FePt : Fe. The simple approach is to use multi layers of Fex/Pt
Fe x /Pt y , where x and
y denote the thickness of each layer. The thicknesses x and y can be adjusted
to obtain a designed volume fraction of FePt phase and Fe phase. The second
approach is to deposit FePt and Fe layers with the consideration of easily
forming the desired Pt-Fe phase. Electron diffraction study has shown that the
Fe, Pt and the FePt layers in the as-deposited films are fcc (Liu CLiu et al. ,
1999b). Figure 6. 14 compares the electron diffraction patterns of the as-
deposited and heat-treated film. The formation of fcc Fe in the film form is
consistent with other observations that many metals can take the close packed
structure in film form (Chopra et al.,
aI., 1967; Luo, 1968).

100 100 111


III
111
III
90 90
80 80
70 70
60 002 60 200
50 113 50
40 202 40
30 30
20 313 20
10 10
0 0

6. 14 Electron diffraction patterns of (a) as-deposited FePt : Fe films and (b)


Figure 6.14
heat-treated FePt : Fe films.

Heat treatment is needed to form the L 100 structure from the as-deposited
film. However, the heat treatment also promotes grain growth. A minimum
annealing temperature of 350 'c is needed to form the L 10 phase for coercivity
CLuo and Sellmyer, 1995). The coercivity of the FePt film with a
development (Luo
Nanostructured Exchange-Coupled Magnets 203

nearly equal-atomic composition shows a monotonic increase from 0.5 to 2.0 T


with increasing annealing temperature from 300 to 700 'c (Liu et al. , 1997a).
The grain size varies from 15 to 100 nm as measured from transmission
electron microscopy. It was found that the thickness of the film has a
pronounced effect on coercivity. The maximum coercivity appears at the
thickness of 35 nm for heat treatment at 300 'c for 15 min.
The investigations started out by synthesizing the FePt/Fe to enhance the
magnetization. The first heat treatment was done at 700 'c for 40 min to
achieve high coercivity. The Pt content was reduced from 50 at. % to 33 at. %
for enhanced magnetization. When Fe content was increased further,
nucleation of magnetization reversal under opposite field started at the soft
phase and a shoulder appeared in the demagnetization curve. In terms of Eqs. (6.8)
and (6. 9), this shoulder is caused by soft regions whose size L is large.
Figure 6. 15 shows the evolution of the microstructure with heat treatment. As
the heat treatment time is reduced, finer grain size of the soft phase Fe is
obtained. Therefore, heat treatment not only helps the L 10 phase to form for

(a)

(b)

(c)

Figure 6. 15 TEM images of (a) (Fe20


CFe20 A/Pt13 16, annealed at 700 'c for 0.5 h(b)
A/PtI3 A) x 16. hCb)
(Fe21 A/PtI2 A) x 16. annealed at 700 'c for 0.5 h. (c) (Fe21 A/PtI2
CFe21 A/Pt12 A!Pt12 A) x 16. heat
treatment : multi-step at 500 'c .
204 W. Liu et al.

high coercivity but also causes the Fe grains to grow. To avoid this dilemma,
a precise heat treatment is necessary, to help form the L 10 phase, but not to
cause excessive growth of Fe grains. Laser surface treatment would be an
excellent means to achieve this. An alternate method is also employed,
consisting of rapid thermal annealing with a high-power lamp that can achieve
a heating rate of 200 C / s. The heating time can be controlled to about one
second. The rapid heat treatment resulted in an extremely fine microstructure,
as shown in Fig. 6. 15c. The two extreme cases of the hysteresis loops
corresponding to Fig.6.15a
Fig. 6. 15a and c are shown in Fig.6.
Fig. 6. 16a,b.
2.0
1.5
~
1.0
S'
:>
E
.,." 0.5
v
'"b
0.0
x
i~ -0.5
--D.5
-1.0
-1.5
-2.0
--{i0
-60 -40 -20 0 20 40 60
H (kOe)
(a)
1500
--
00-- Parallel
1000 - - Perpendicular

~
u 500
-3~E
E 0
~
~
-500

-1000

-1500
--{i0
---{)O -40 -20 0 20 40 60
H(kOe)
H (kOe)
(b)

Figure 6. 16 Hysteresis loops of FePt : Fe films. (a) Corresponding to the specimen in


Fig.5. 15a,b
Fig. 6. 15a, b corresponding to the specimen in Fig.5.
Fig. 6. 15c.

Liu et al. (1998a) sputtered Fe layers having twice the thickness of the Pt
multi layers with an atomic Fe-to-Pt ratio of about 2 : 1. To
layers to prepare multilayers
avoid excessive grain growth, a special rapid thermal annealing process that
includes more than one step of treatment was adopted. The first step was
annealing
anneal ing at 500C for 5 s, with a heating rate of about 200C per second.
The second step involved a 15 min annealing
anneal ing in a furnace at about 450 C .
Nanostructured Exchange-Coupled Magnets 205

With this process it has been possible to keep the grain size of the second
phase below 10 nm as shown in Fig. 6. 15c, obtaining coercivities higher than
1.0 T; and a maximum energy product higher than 318.4 kJ/m 3 measured with
the field parallel to the sample can be achieved. An extremely large energy
product, 420.3 kJ/m 3 (52.5 MGOe) , corrected by an effective demagnetizing
factor in the perpendicular direction, was achieved at room temperature as
shown in Fig. 6. 16b. Processing nanostructure and magnetic properties of
FePt : Fe films are concluded in Table 6. 2. A nanocomposite with desired
nano-structure consisting of the FePt and a soft Fe-rich phase was formed.
Thus, this system constitutes an excellent prototype for intergrain exchange
coupling and for developing very high-energy products.

Table 6.2 Processing,


Processing. nanostructure and magnetic properties of FePt : Fe films (Liu
CLiu et al. ,
1997a. 1998a. 1999b).
Processing Microstructure Properties
CInitial film configuration,
(Initial configuration. Cgrain size of soft phase,
(grain phase. CIntrinsic coercivity.
(Intrinsic
No.
composition,
nominal composition. grain size of hard phase.
phase, polarization. energy
heat treatment) volume fraction of soft phase) product)

CFe20/Pt13)
(Fe20/Pt13) x 16 A.
A, not observed,
observed. 1.91 T,
T.
1 Fe-39. 4 at. % Pt. 50-200 nm,
50-200nm. 1.29 T.
700'0
700'C 0.5 h 0 236 kJ/m 3

CFe21/Pt12)
(Fe21/Pt12) x 16 A.
A, 5-20 nm,
nm. 1.56 T,
T.
2 Fe-36.
Fe-36.33 at. % Pt. 30-150
30- 150 nm. 1.41 T,
T.
700'0
700'C 0.5 h 4% 307 kJ/m 3

CFe21/Pt12) x 16 A.
(Fe21/Pt12) 3-8 nm,
nm. 185 T.
4 Fe-36.
Fe-36.33 at. % Pt. 50-100
50- 100 nm. 1.63 T,
163 T.
Multi-step at 500 '0
'C 4% 420 kJ/m 3

Figure 6. 17a shows a high-resolution transmission electron microscopy


500 'c shown in
image of a FePt : Fe film after multi-step heat-treatment at 500'C
Fig. 6. 15c. The small grain size of the soft phase ensures the exchange
coupling. Figure 6. 17b and c are energy dispersive X-ray spectra taken from
hard phase and the soft phase, respectively. The compositions deduced from
the spectra are 31 at. % Pt for the soft phase and 35 at. % Pt for the hard
phase. Based on above TEM studies, there are two issues that can lead to a
reduction of the energy product of the film. First, ideally the soft phase should
be the pure iron which has a very high magnetization, but Pt reduces the
magnetization of the soft phase and therefore the magnetization of the film.
Second, the stoichiometric compound FePt has the highest anisotropy and
How~ver in the nanocomposite FePt : Fe, the Fe
highest coercivity. Howl?ver Fe,x Pt 1 -- ,x
phase contains a higher Fe content that reduces the anisotropy and coercivity.
The dilution of the soft phase and hard phase is the result of phase equilibrium.
This suggests that the best nanostructure must be realized by non-equilibrium
synthesis.
206 W. Liu et al.

(a)
PI
Fe
b
t

(b)

Fe
c
t
c:E::l PI
Pt
o
tl1$
-<
<l:
Fe

Energy
(e)
(c)

Figure 6. 17 High-resolution transmission electron microscopy image of a FePt : Fe film


shown in Fig. 6. 15c. Figures 6. 15b and 6. 15c are energy dispersive X-ray spectra taken
Fig.6.
from hard phase and the soft phase, respectively.

Based on above consideration, nanostructured design and non-equilibrium


synthesis were considered to minimize the dilution of the soft and hard phases.
The initial film structure was designed as: Fe (x) /FePt (y) where x and y
denote the thickness of each layer. This film structure with the hard phase
layer FePt ensures that minimum diffusion time is required to achieve the L 10
phase. Figure 6.18
6. 18 compares the hysteresis loops of Fe : FePt and Fe : Pt.
Both films have the same content of Fe. The magnetization reversal occurs in
Nanostructured Exchange-Coupled Magnets 207

much narrower regime in Fe : FePt than in Fe : Pt. This suggests that the L 10
phase in Fe : FePt specimen has more uniform composition, while the L 10
phase in Fe : Pt specimen has larger composition range with different
anisotropy and thereby decreasing the coercivity. This phenomenon can be
well explained by the initial film layer structure. The L 10 phase in Fe : FePt
has much more uniform composition and anisotropy while the L 10 phase in
Fe : Pt has much larger composition range and anisotropy because of longer
diffusion range required for the formation of L 10 FePt.
~emu
M 250 Ilemu

H
~---'-/-~-+-~--f-'--1~5kOe
~-'-I-~-+-~+'--1~5

(a)

~emu
M 250 Ilemu

H
~-4-~-+~~'-'--:-1-=-'5 kOe

(b)

Figure 6.18 Hysteresis loop of (a) (Fe12 A/(FeP040


A/(FePt)40 A) x 10, and (b) (Fe16 A!Ptl0
A/Ptl0
A) x 16. Heat treatment for the two cases were at 450"C
450'C for 5 s.

The thickness of Fe should be as thick as possible for remanence


enhancement without destroying the coercivity. The thinner the Fe layer, the
easier it is to become diluted. Figure 6. 19 compares the effect of layer
thickness of Fe keeping the content of Fe constant. The heat treatment was
done at 450 'c for three seconds. As the Fe layer becomes thick, nucleation of
reverse domain occurs more easily in the soft phase, destroying the
coercivity.

6.3.2 Rare-Earth Cobalt Magnets


6.3.2.1
6. 3. 2. 1 Phase Diagrams and Crystal Structures

Although many compounds between rare earth and Co, such as RCoz , RC0 3 ,
Rz COl'
C0 7 , RCo s and Rz CO I7 , have been found in the R-Co system, only RCo s and
COIl'
208 W. Liu et al.

250llemu
250 Ilemll

20 kOe
20kOe

(a)
250llemu
250 Ilemll

20 kOe

(b)

250llemu
250 Ilemll
_-::::>""-
~-::::--

20 kOe

(c)

Figure 6.19 Hysteresis loops of (a) (Fe9 A/(FePt)30


A/(FeP030 A) x 14, (b) (Fe12 A/(FePO
A/(FePt)
40 A) x 10, and (c) (Fe18 A/(FeP060 A) x 7.

COp- (R = Sm) based alloys have been sucessfully fabricated for permanent
R2 Cow
magnets. The RCos compounds crystallize
crystall ize in the hexagonal CaCus -type
CaCus-type
structure with the space group P6/mmm. Figure 6.20 shows the CaCus-type
structure. In a unit cell, the Ca (or R) atom occupies the 1a site and Cu (or
Co) atoms occupy 2c and 3g sites. The CaCus-type
CaCus -type structure constitutes a
basic structure of R-T compounds, from which many structures of the R-T
compounds can be derived (Bouchet et al. al.,, 1966; Buschow, 1971).
Figures 6. 21 and 6. 22 show the equilibrium
equil ibrium phase diagrams of Sm-Co
(Massalski, 1990c) and Pr-Co (Massalski, 1990b) alloys, respectively. The
detailed crystal-structure data of Sm-Co (Massalski, 1990c) and Pr-Co
( Massalski, 1990c) alloys are summarized in Tables 6. 3 and 6. 4
respectively. The R-Co diagrams with other rare earths are generally similar,
but with some minor, systematic variations. The RCo s phases are generally
unstable at room temperature, but can be retained as metastable phases at
room temperature by rapid cooling (Strnat, 1972).
Nanostructured Exchange-Coupled Magnets 209

eR OTM(g) TM(c)
0TM(c)

Figure 6.20 Schematic diagram of the unit cell of the hexagonal CaCus-type structure
with space group P6/mmm formed by RCos compounds.

Table 6.3 Co-Sm Crystal Structure Data (Massalski, 1990c).


Composition, Pearson Space Strukturbericht
Phase Prototype
Sm (at. %) Symbol Group Designation
(a-Co)
(ex-Co) ao to -1.
-1.55 CF4 Fm 3m
3m Al Cu
(e-Co) -0 HP2 P6 3/mmc A3 Mg
~-Co17Sm2
~-CoI7Sm2 -10.5 HP38 P6 3/mmc Ni 17 Th2
ex-Co 17 Sm2
a-Co -10.5 HR19 R3m Zn17 Th2
COs+,
+ x Sm -16 - 17
CO s-, Sm
Cos-xSm -17 - 18
Co 19 Sm
CO'9 SmsS -20.8 HR24 R3m Co 19 Ces
CO'9 CeS
HP48 P6 3/mmc Ce2 C07 Cu3
Ce2C07Cu3
C07
7 Sm2 -22.2 HR18 R3m C07
7 Er2
Er2
HP36 P6 3/mmc Ce2 Nb
Ce2Nh
C03Sm 25 HR12 R3m Be3Nb
C02Sm 33.3 HR4 R3m C02Sm
CF24 Fd3 m C15 CU2Mg
CU2 Mg
C0 4 Smg
Co,Sm9 -69.2 0* *
CoSm3 75 OP16 Pnma DOli
DO" Fe 3C
Fe3C
(y-Sm) -100 CI2 Im3m
1m 3m A2 W
(~-Sm) -100 HP2 P6 3/mmc A3 Mg
(ex-Sm)
(a-Sm) -100 hR3 R3m a-Sm
ex-Sm

Other reported phases

CosSm -16.7 HP6 P6/mmm D2d CaCus


HP'
C02Sms -71.4 MC28 C2/c MnS C2
210 W. Liu et al.

Table 6.4 Co-Pr Crystal Structure Data (Massalski, 1990b).


Composition, Pearson Space Strukturbericht
Phase Prototype
Pr
Pr(at.%)
(at. %) Symbol Group Designation
(ex-Co)
(a-Co) -0 CF4 Fm 3 m Al Cu
(e-Co) -0 HP2 P6
P6,/mmc
2/mmc A3 Mg
C0 17 Pr2
CO ,7 10.5 HR19 R3m Zn,7
Zn17 Th2
C0
Cos5Pr 16.7 HP6 P6/mmm D2d CaCus
CaCu5
l9 Pr
CO ,9 Pr5S 20.8 HR24 R3m C0 ,9 Ce
CO 19 Ce5S

(3-C0 7
13-C0 7 Pr2 222
22.2 HR18 R3m 7 Er2
C0 7 Er2
a-C0
ex-C0 7
7 Pr2 22.2
222 HP36 P6 3/mmc Ce2Nb
Ce2Nh
C03Pr 25 HR12 R3m Be3Nb
C02Pr CF24 Fd 3 m
Fd3 C15 CU2Mg
C01.7 Pr2 -54.1 hP"
CO2Prs
Pr5 -71.4 MC28 C2/0 MnS5C,
C2
CoPr3 75 OP16 Pnma DOlI
DO'I Fe3C
(13-Pr)
(3-Pr) -100 CI2 13m A2 W
(ex-Pr)
(a-Pr) -100 HP4 P6 3/mmc A3' ala
exla

Weight percent samarium


0 10 20 30 40 50 60 70 80 90 100
1600

,
1495'C
1495C

1400
,",
'I
1---- 1260'C
1240'C L
1200
1200 :I 1200'C
I 13Co 7Sm 2
I 1074'C
I
:J 1000
:..-
'--' t---.
t----
I (aCo)
.,~
~
::J
:::l

.... 800 I E
'"0-
<U
0- I "1-
'1-
E E E
E I 0
VJ
'1-
0-
0 VJ
C/l ,,:!-,
C/l
~
.<U
:- u
U
~d
U a 6'
0' 0
d
<:j u
U u
600
422'C
422C

400
II E'"
E E
If
VJ
C/l
II VJ C/l
(ECO) II
(eCo) 00- 0 (aSm)
u u
200 I I
I
I I I
I I I
0 10 20 30 40 50 60 70 80 90 100
Co Atomic percent samarium Sm
Figure 6.21
6. 21 Phase diagram of the Co-Sm system (Massalski, 1990c).
Nanostructured Exchange-Coupled Magnets 211

Weight percent praseodymium


1600
o 10 ZO
20 30 40 50 60
r--'-----'---'-----'----'--~---'----~-----'-----___1
r--'----'-----'-~-~-----'----'----.L..-----'------___j
70 SO 90 100

1495"C
149S'C

L
I ZOO
1200
1125 "C
112S'C
11IS 'C
I1\S "C
1000 1057 C
f.J 931"C
931'C 931"C
931'C
'--'
~
::l
''"
SOO
e<l)
(aCo)
(nCo)
"
0-
0-
E
~
59S"C
S9S'C
570"C
600 .:'
.s-' S70'C
5S0"C
5S0'C S41'C
~
Q"
0- 66
4ZZ"C
422'C U
U 55S"c'I
S5S'c'
400 1I
.:'
.s-'
.:'
.s-' ~... rf
.::'
A.. ... ... A..
Q" ~I
~I
M

~I
A..
~I 0;;,1 ~II 10;:.,
10;:, ";:.1 ::1
:::1 ";:.10::1I
'7,10::
ZOO
200 U
U1 UIU 01 0Il 1
111 UU U U1 U1
U
01 01
11 (aPr)
1I II I I 1I U1U 1I
UIU
(ECO) 1I
(eCo) I II I I I 1I I
II I II I I I I I
0
0 10 ZO
20 30 40 50 60 70 SO 90 100
Co Atomic percent praseodymium Pr
Figure 6,22
6" 22 1990b) .
Phase diagram of the Co-Pr system CMassalski, 1990b).

Strnat (1972) summarized the binary R-Co phases stable at room


temperature, and metastable phases formed by quenching from a higher
temperature as shown in Fig. 6.23. For each idealized phase composition the
crystal symmetry is listed (in some cases for room- and high- temperature
structures), and the bars indicate the possible phases formed with various rare
earth elements. For magnets, only those in the upper half of the bar graph are
of interest: a 2 : 17 or 1 : 5 as the principal flux-producing phase, the others,
down to 1 : 3, as secondary minor phases that influence the coercive force and
its temperature dependence. The compounds are ferro- or ferrimagnetic above
room temperature. The two meandering dashed lines define the regions where
the Curie point is above 20 and 400 'c , respectively. The Curie Temperature
SmCo2 , Sm9 C0 4 and Sm3Co,
of SmC0 PrCo2 , Pr2 Co1.7'
Sm3 Co, PrC0 COl 7' Prs
Pr5 CO2 and Pr3Co
Pr3 Co phases
are below 20 'c, therefore, these compounds are not considered as
candidates for permanent-magnet materials. The Curie temperatures of RC0 RCo 3 ,
= Sm and Pr) are between 20 and 400 'c , and those of
R2C0 7 and Rs5C0 19 (R =
RCos5 and R2 C0 17 are above 600 'c (R = Sm and Pr). Liu et al. (1998d)
reported a new metastable phase SmC0SmCo3 that is formed upon annealing the film
containing Co-22at. % Sm, at 500 'C. This phase has the 00 19 structure in
which the Sm atoms take ordered positions of a triangular pattern in the close-
packed planes. Figure 6.24 shows the 00 19 structure. A very high coercivity of
2. 9 T was obtained from this phase. The lattice parameters deduced from this
212 W. Liu et al.

phase are a = 0.256 nm, C = 0.419 nm compared to the lattice parameters of


Co 0=
a = 0.2505 nm, C = = 0.4065 nm. The lattice parameters between the 00 19
phase and the Co phase are close to each other, suggesting a low energy state
at the phase interface between Co and DO
00 19
19., Therefore, two-phase structures

composed of Co and 00 19 phases are candidates for exchange-coupled


magnets.

Symmetry
~
I-J
0 RCo lJ Cubic
0
(II
(I) Hex
X
'T
/\ R2 CO l7
Rhomb
~
K c--,
(2)

1i RCo 5s
-
W..#'-E-E&
WYfff&
r----------
Hex
(J)
(3) Hex
~
I-J R SCO l9 I
0 I Rhomb
N
Rhomb
2C0 7
/\
1\ R2
~
K Hex

t
RCoJJ Rhomb
----- -
RCo2 Cubic
~
I-J L_J
0
N
R2Co JJ ~ Ol1horhomb
Orthorhomb
v
~
K R2CO l7 ~ Hex
1::>i R4Co]
Co J
:::l Hex
ec..
[!
Rp2 CO(4)
Co(4)
'"
Q)
Q.
F2
E
E
.~
Q)
.c;
R24 CO
C0 11II ~ Hex
::>
:::l
u R9 C044 Orthorhomb
(5)
R]C6
RJCo Orthorhomb

6. 23 A summary of rare earth-cobalt intermetallic phases (Strnat, 1972). Notes:


Figure 6.23
( 1) Where double bars are drawn, the upper bar indicates the crystal symmetry and
existence range of a high-temperature modification; the lower bar represents the structure
reported to be stable at room temperature,
temperature. (2) For the heavy rare earth, Tb and Dy
through Tm, the phase with the CaCus -type structure exists only as RCo
RCo5+
s+,x with Co-excess
that increases with increasing atomic number of R. The RCos phases are generally
unstable at room temperature, but can be retained metastable at room temperature by
rapid cooling. (3) Additional 5 - 19 phases have been reported, but their existence-
certainly as a stable room-temperature phase - is questionable.

6.3. 2. 2 Magnetic Properties of R-Co and Related Magnets


Strnat summarized basic magnetic properties of the R-Co compounds for
permanent magnet applications (Strnat, 1988). The results are given in
Table 6.5.
Table6.5.
Table 6.5 shows that RCo s (R = Ce, Pr, Nd, Sm or Y) compounds exhibit
RCos(R=Ce,
fairly good overall magnetic properties, such as sufficiently high Curie point,
very high magnetocrystall ine anisotropy and high saturation magnetization.
magnetocrystalline
Within the RCoss series, SmCoss has attracted most interest
Nanostructured Exchange-Coupled Magnets 213

-.
--.- --.-
--e_

Sm



Figure 6. 24 The 00,9
00 19 structure.

Table 6.5 Basic magnetic properties of compounds of interest for R-Co magnets (room
temperature values, except T ee )) ,, based on the data given by Strnat (1988) for single-
crystals and powder samples.
B,(T)
B,CT) K 1 (10 6 J/
K, Theoretical value.
Compound TaCC) /.10 H a (n[IJ
1.10 (n[']
(1.10 M,)
(/.10M,) m33)[1]
m )[lJ (BH) max (kJ/m3) [2]
[2J
YCOSs
YCo 1. 06
1.06 630 5.5 13.0 223
LaCo s 0.91 567 6.3 17.5 164
CeCos[3]
CeCos[3J 0.77 380 6.4 21. 0
21.0 117
PrCos 1.20
1. 20 620 8.1 17.0 286
NdCos 1. 22
1.22 637 0.24 5.0 295
SmCos[4]
SmCos[4J 1.14 727 11-20 25.0-44.0
25.0- 44. 0 259
Sm2 Co 17 1.25
1. 25 920 3.2 6.5 310
Sm2 (Coo.
(Coo 7 Feu)
Fe03) 17
'7 1. 45
1.45 840 3.0 5.2 417
Sm(CoO.87 C UO.13
CUO. 13 ),.8
)7.8 1. 09
1.09 847 3.3 7.7 237.6
[1] K I calculated from H aa =
[1 J In most cases H a was measured as the hard-axis saturation field, and K, =
2K,/lJoM,.
2K1/J.loM,.
[2J The limiting energy product values were calculated as lJoMU4.
J.loMU4.
[3J A wide range of B, and T e values has been reported,
reported. possibly due to impure cerium. Since the
most likely R-impurities (La,
(La. Pr and Nd) increase M, and T e , the lowest reported values are listed.
[4J For SmGos, a wide range of property values is found in the literature, especially for the
anisotropy (H a , K 1).
I). Great discrepancies may originate from the fact that available fields in most

Ha and a long extrapolation results in inaccuracy of measurements. The


experiments are much lower than H
inaccuracy also originates from the use of powders instead of single crystals. In addition,
addition. off-stoichiometry
in "SmGos" (as SmGoS+6)'
SmGos+ 6)' due to quenching from higher temperatures, results in significant variation of
H a with O.
214 W. Liu et al.

because of its extraordinarily large anisotropy (Franse et al., aI., 1993); and
SmC0
SmCo5-based permanent magnets have been produced for decades. Although
the compound PrCoPrC05 has an unfavorable anisotropy (easy cone) at room
temperature, its saturation magnetization exceeds that of SmC0 SmCo5 (Strnat,
1967>.
1967). Not long after the magnets based on the 1 : 5 compounds were
developed, the next generation of magnets based on the 2 : 17 phase
containing more Co also joined the ranks of rare-earth permanent magnets,
that provide higher saturation magnetization and higher Curie temperature than
those based on the 1 : 5 compounds. The real structure of the 2 : 17-based
magnets are multi-phase metallurgical systems with complex microstructures,
generally not in an equil ibrium state, and they always contain more than just
equilibrium
two elements. Their coercivity is caused by domain-wall pinning, as compared
to nucleation type magnets (Zijlstra, 1970; Skomski and Coey, 1999). With
an optimization of the composition and sintering heat treatment, the energy
product record was raised to 263 kJ/m 3 , combined with JJo Po He = 1.3 Tim for
Sm(Co,Fe,Cu,Zr)767
Sm( Co, Fe, Cu, Zr)767 (Mishra et al. , 1981). The 2 : 17 based magnets have
a complex three-phase structure. TEM observation showed that the
nanostructure of the magnet consisted of rhombohedral cells of the 2 : 17
phase with Th zZn17 structure and the cells were surrounded by the 1 : 5 phase
Th2Zn17
with CaCu5 structure. The size of the cells is about 100 - 200 nm, and the
width of the cell wall consisting of 1 : 5 phase is about 5 - 20 nm. Zr rich
precipitate layers parallel to the basal plane penetrate the cells and are
coherent with the 2 : 17 phase lattice. Each alloying element plays a role in
stabilizing the structure. The Fe-containing 2 : 17 phase contributes to a high
magnetization and the 1 : 5 phase contributes to high coercivity, and Cu
promote precipitate to form and Zr is the major precipitate former. However,
magnets made of PrC0PrCo5 have never been commercialized. One reason for this
is unfavorable phase equilibrium since undesirable soft magnetic phases exist
on both sides of the target hard magnetic PrCo PrC05 phase. Only low levels of
coercivity (less than O. 6 T) have been achieved in sintered or melt spun
binary Pr-Co alloys (Furest et al. , 1993; Velu et al. , 1989). Recently, it has
been shown that carbon additions result in great enhancement of the coercivity
of Pr-Co alloys (Furest et al., 1993; Lewis et al., 1996). Branagan et al.
800 C , the quaternary Pr-Co-Ti-C
(2000) reported that after heat-treating at 800C,
alloy developed high coercivity ( 1. 67 T) and an energy-product of
1.67
kJ/m
68.5 kJI m3 . Chen et al. , (1999a) obtained the highest coercivity of 2.
2.37
37 T in
nanostructured PrCo
PrC05 based powders produced by mechanical milling and a
subsequent annealing. Overall properties with a high coercivity of 1.63 T
along with a high Mrl M M s ratio of O. 66 and a moderate maximum-energy
product of 92.3 kJ/m 3 have been developed in stoichiometric PrCo PrC05 powders
prepared by milling for 4 h and annealing at 800C
800C for 1 min. Because of the
Pr2Co
planar anisotropy of Prz l7 (Schaller et al., 1972; Ray et ai.,
CO 17 aI., 1972), the
Pr-Co system was not considered as a candidate for precipitation-harden
magnets. Recently, hard magnetic properties have been found in homogenized
Nanostructured Exchange-Coupled Magnets 215

and subsequently aged Pr-Co-Zr alloys consisting of a Pr2 (CoZr) 17 matrix with
(Pr, Zr) Cos precipitates formed after aging the homogenized alloys with the
(Pr,
(Pr,Zr)CoS+
Zr) Co s+ 66 structure. Anisotropic Prl1.5Zr4COSS
Prll.S Zr4 Coss powders with coercivity of
product of 57. 3 kJ/ m3 were obtained (Gabay et al. , 2000).
O. 41 T and energy proooct
Since the permanent-magnet potential of the intermetallic compound
SmCos was found (Strnat, 1967), a var variety
iety of production methods have been
developed, such as compressing (Buschow et al., ai., 1968, 1969), sintering of
magnetically aligned powders (Das, (Oas, 1969), melt-spinning (Croat, 1982),
sputtering (Cadieu et al. , 1985), mechanical alloying (Wecker et al. , 1991)
and pulsed laser deposition (Cadieu et al., 1998). Foner et al. (1972)
obtained a room-temperature energy product of 191 kJ/m 3 for a material
63. 7 wt. % Co, 32.6 wt. % Sm and 3.3 wt. % Sm2 0 3 , The
containing 63.7
coercivity of 6.6 T in SmCos synthesized by means of chemical reduction
during mechanical alloying was reported for a bulk Sm-Co systems at room
temperature as shown in Fig. 6. 25 (Liu et al., 1992). The highest value of
/JOM He equal to 7.5 T has been obtained in this laboratory. Liu et al. , (1998d)
JJOM
reported a coercivity of 4.5 T in sputtered Sm-Co film with the CaCus structure
annealed at 600 'c .
SO
50

25

-25

-50':-:-----:"':-:-_:-:-_-:-----:'":-----:-:-:-____=_'_
-50L-_-'--_-'--_-'--_-'--_-'-----'
-ISO -\00
-\50 -100 -so
-50 0 so
50 100 ISO
150
H(kOe)
Figure 6. 25 Hysteresis loop for an Sm-Co
8m-Co sample annealed at 973 K.

Sintered magnets based on Sm2 (Co, Cu, Fe, Zr) 17 exhibit outstanding
(JJOM He >
magnetic properties and still provide the highest coercive fields (/JOM
3.5 T at room temperature) as well as Curie temperature (T c > 1000 K)
attainable today among all technical rare-earth permanent magnets (Buschow,
1997). Popov et al. (1990) reported that Sm2 CO l7 -based bulk-hardened
magnets exhibited very diverse temperature dependencies of coercivity
M He (T), including one with a positive temperature coefficient. This
"anomalous" M He (T) has attracted much attention in recent years (Liu et al. ,
1998c; Goll et al. , 2000; Zhou et al. , 2000; and Liu et al. , 2000b), because
of the need for developing magnets for high-temperature applications. The
phenomenon appears to be nearly universal for all Sm-Co magnets having a
microstructure consisting of Sm2 C01~ cell surrounded by the SmCos phase
(Gabay et al. , 2001)
2001),, and a certain amount of additives (Zhou et al. , 2000).
216 W. Liu et al.

According to the above results, there is no doubt that PrC0


PrCo5s and, in particular,
8mCos5 based phases should constitute good candidates for the hard phase in
8mC0
exchange-coupled magnets.

6.3. 2. 3
6.3.2.3 Artificial Two-Phase Systems
Chen et al. (2000a) investigated nanocomposite PrCo PrC0s5 : Prz
Pr2 COil
COll powders
with a different volume fraction of Pr2 CO I7 . They were synthesized by
Prz COl?'
mechanical milling the mixtures of PrC0 PrCo5s and Prz COil
COl? alloy powders, and the
highest maximum energy product of 94 kJ/m 3 was obtained in a mixture
consisting of 70% 70 % PrC0
PrCo5 30 % Prz COil'
s and 30% Co ll They also achieved a maximum
energy product of about 93. 9 kJ/m 3 in mechanically milled nanogranular
Pr x Co - x powders with x = 18. 0 (Chen et al., 2000b). Microstructural
IOO -
CO 100
studies revealed that a uniform nano-scale PrCo PrC0s5 : PrzCo lll? microstructure with
an average grain size of about 20 nm is developed in these PrC05/PrZCa7 PrCos /Prz Cal
powders. Evidence of intergranular exchange coupling is found in the mixture
of PrC0
PrCo5 Przz COil
s and Pr COl? powders, indicating that the enhanced energy product is
due to the exchange coupling between the magnetically hard PrCo PrC0s5 and soft
Pr ZCo ll
PrZCo 17 nanograins.
Recently the multilayer thin films of C0 Co 3 8m/Co and C0 3 Pr/Co have been
synthesized by sputtering. The as-deposited nanostructures in the Co C0 3 8m : Co
and C0 3 Pr : Co systems are a mixture of amorphous phase and crystallites of
about 5 nm. The volume fraction of the crystallite phase varies with the
content of 8m and Ar pressure during deposition, as shown by Liu et al.
((1994).
1994). Low Ar pressure and low 8m content produces low volume fraction of
amorphous phase. Pr has a strong tendency to stabilize the amorphous phase.
High coercivity can only be achieved by post-annealing. Figure 6.26 shows the
TEM image and the hysteresis loop of the C0 3 Pr : Co film. The grain sizes of
both the hard phase and soft phase after heat treatment at 500 500'C'c for 30 min
are around 10 nm. 8uch a grain size is at the threshold of effective exchange
coupling for the hard phase and the soft phase with a slightly visible shoulder in
the hysteresis loop shown in Fig 6. 6.26b.
26b. In both C0 Co 3 8m:
8m : Co and C03 Pr:Pr : Co,
the hard phase has been identified as the D0 00 19 structure C0
Co 3 8m or C0 3 Pr.
Figure 6.27 compares the electron-diffraction patterns of the soft phase Co and
the hard phase C0 3 Pr. The extra Iines lines due to ordering are indicated by the
arrows. Figure 6.28 shows energy dispersive X-ray (EDX) (EOX) spectra confirming
the composition of the hard phase and the soft phase. The Cr is from the
coating layer of the film. It is interesting to notice that the soft phase is not
diluted by the heat treatment as in the case of Fe in FePt : Fe.
In both C0 Co 3 8m : Cox and C0
Co 3 8m : Pr x systems, multi-step heat treatment
is most effective for achieving high coercivity and exchange coupl ing. The first
step is to use a high power lamp furnace with a heating rate of 200 200'C'c to heat
500 'c for 20 s and then 500'C
treat at 500'C 500 'c for 30 min in a vacuum furnace (Liu
et al. , 1999b). The coercivity higher than 4.2 T was found for C0 Co 3 8m : Cox.
Cox'
As the thickness of Co layer increases the coercivity decreases and the
Nanostructu red Exchange-Coupled Magnets
Nanostructured 217

(a)

6
4
2
0
-2
-4
-6
-fi
-60
-fi0 -40 -20 0 20 40 60
kOe
(b)
Figure 6. 26 (a) TEM images of C03 Pr : Co, (b) hysteresis loop.

Figure 6. 27 Comparison of electron diffraction patterns of (a) Co and (b) heat treated
C03 Pr : Co film.

Cox system , /JOM


magnetization increases. In the C0 33 Pr : Coxsystem, Hec of 1.0 T and /Jo
IJoMH IJo M s
of 0.83 T for x = 1 nm were obtained. As x increased to 8 nm, nm. coercivity
dropped to 0.50 T but magnetization /Jo IJo M s increased to 1. 21 T.
1.21
Figure 6. 29 shows hysteresis loops of a PrCo35 : Co nanocomposite
sample measured at different temperature (Liu et al., al.. 2000a).
2000a) . The
squareness of the loops decreases with decreasing temperature,
temperature. from 0.89O. 89 at
300 K to 0.51
O. 51 at 5 K. Small squareness is a direct sign of ineffective intergrain
exchange coupling in the nanocomposites. The reduced squareness is an
indication of a nucleation or switching field distribution; that is,
is. the grains are
not exchange-coupled. Such change can be explained by the enhancement of
218 W. Liu et al.

2 2
Co
~ Co
6
X X
~ ~

'"c
tl
C
:::l
'"
E
:::l
0 Cr o
U U

Co

0
4 6 8 10
eV eV
(a) (b)

Figure 6. 28 Energy dispersive X-ray ( EDX) spectra of PrCo 35 , (a) taken from soft
Cb) the hard phase.
phase and (b)

anisotropy at lower temperatures as calculated by Goll et al. (1998) and


Hadjipanayis (1999).
Magnetic hysteresis and intergrain exchange coupling in nanostructured
PrCo : Co composite films prepared by multilayer sputtering and subsequent
anneal
annealinging were investigated by Liu et al. (1997b, 1998b, 2000a), and the
coercivity is related to film morphblogy, morphology, especially the Co phase fraction.
Evidence for the hard-soft-phase exchange coupling has been found.
In bulk systems, mechanical alloying or milling has been mainly applied to
prepare exchange-coupled magnets based on SmC0 SmCo5 s . Ding et al. (1993b)
3
3
reported a maximum energy product of 143.3 kJ/m in optimally treated Sml2 Sm12 5
5

C08U
875 powder. A coercivity of 5.7 T was obtained in mechanically alloyed
Sml9C081
Sm19Co81 after heat treatment at 800C 800C for 30 min. (Ding et aI., 1993c). A
remanence enhancement was achieved in mechanically alloyed and heat-
Sml3 (Co'-xFe
treated Sm13(Co 1- xFe x )87 powders containing a two phase mixture of Sm(Co,
Fe)5
Fe)s or Sm2(Co,Fe)7 and bcc (Fe,Co). The coercivity of 0.40T to 0.50T
and the maximum energy product of 159.2 kJ/ m3 3
were measured for this two-
phase system after applying a demagnetizing-factor correction (Ding et al. ,
1994). They also obtained a maximum energy product of 127.4 kJ/ m3
kJ/m 3
in
mechanically alloyed Sm-( Co, Fe) permanent magnets annealed at 700C
Sm-(Co, 700C for
10 min (Smith et aI., al., 1995). The anisotropy field, between 20 and 30 T at
room temperature, was estimated from approach to saturation. This indicates
the presence of Sm2 (Fe, CO)7 (J,Jo Ha~
(/10 H a ""'" 30 T at room temperature) or Sm(Co,
Fe)5
Fe)s (J,Jo
(/10 H Ha~
a ""'" 30 T at room temperature) rather than the presence of Sm2 (Fe,
Co) 17 ( J,Jo H aa ~
17 (/10 ""'" 10 T at room temperature) in exchange-coupled Sm-Fe-Co

magnets after mechanical alloying and subsequent annealing (Dahlgren et al. ,


1998). Zhang et al. (2001a) (2001 a) reported that depending on the Sm content in as-
mi lied Sm
milled SmxCox C0 5
s (x =
=0.0 . 65-1.
65 - 1. 3) plus 20 wI. wt. % a-Fe powders, the hard phase
is a 1 : 7, 1 : 5 or 2 : 7 phase after a heat treatment at 550C 550C for 30 min. A
Nanostructured Exchange-Coupled Magnets 219

300 K
300K

250 K

200 K
=
<::
0
""
.~
N
.~

=
<::
OJ)
OJ)
co
150 K
~'"
::E
lOOK

50 K

5K

-50 -40 -30 -20 -10 0 10 20 30 40 50


Field (kOe)
Figure 6. 29 Variation of magnetic hysteresis loops of CoPr3
COPr3 : Co with temperature.

coercivity of 0.6 T and a maximum energy product of 141.7 kJ/m 3 were


achieved for x == 1.0.
In multilayer systems, Fullerton et al. (1998a) demonstrated that
epitaxial growth of 8m-Co/Co superlattice films prepared by magnetron
sputtering approach the ideal nanostructure of exchange-spring hard/soft
ferromagnets with aligned
al igned hard-magnet layers. The oriented growth was
achieved by magnetron sputtering onto epitaxial Cr ( 100) -buffered single-
100 )-buffered
crystal MgO( 110) substrates. X-ray diffraction (XRD) and TEM studies show
that the superlattice is structurally coherent, with a high density of twin
boundaries. At room temperature, the 8m-Co films have a uniaxial anisotropy
field of 20 T and a coercivity of more than 3.0 T. The magnetization of the soft
layer is exchange-coupled to the aligned hard-magnetic layer and switches
reversibly as expected for an exchange-spring magnet. Fullerton et al. (1999)
remarked that although a single 8m-Co layer has coercivity as large as 3.0 T,
the low saturation magnetization of 8m2 Co]
Co? results in the maximum magnetic
220 W. Liu et al.

product (BH)
(BH)max 87. 6 kJI m3
max of only 87.6kJ/m
3
. When interleaved with Co layers, the
total saturation magnetization of the multilayer initially increases, and the
( BH ) max increases to about 111. 4 kJI m3 3
.. Liu et al. (1999a) reported a
remarkable increase of the coercivity in the CrOOO A) I[ SmCo35 ( 100 A) I
A) J25 ICr( 100 A) annealed at 500 'c for 20 min, up to 4.0 T. AI-Omari
Co(30 A)]25
and Sellmyer (1995) studied Sm20 COao COeo IFe65 C0 35
35 bilayer and multilayer films
with Cr underlayers and overlayers. For a fixed Co-Sm layer thickness, the
magnetization of these samples increases with increasing FeCo layer
thickness. The coercivity, anisotropy constant and anisotropy field for films
with a fixed SmCo layer thickness were found to decrease with increasing Fe-
Co layer thickness. The maximum energy products for these samples varied
from about 47.8 kJ/m 3 at room temperature to 207 kJ/m 3 3
at 30 K. The
reversible demagnetization curves were measured and found to be consistent
with the behavior expected for exchange-spring magnets.

6.3.3 Nd-Fe-B-Based Magnets


To obtain a nanocomposite permanent magnet with good properties, it is
necessary to have a magnetic component with excellent hard magnetic
properties. In this respect, the Nd2 Fe14 B compound and its modifications
through various element additions are promising. Croat et al. and Sagawa et
al. first reported Nd-Fe-B permanent magnets prepared by melt spinning and
by a conventional sintering process, respectively, at the 1983 Conference on
Magnetism & & Magnetic Materials held in Pittsburgh (Croat et al., 1984;
Sagawa et ai.,
al., 1984). Precursor work includes researches on binary Nd-Fe
alloys and on B additions to R Fe-alloys prepared by melt spinning by Croat
( 1981, 1982) and by Koon and Das (1981), respectively.
Sagawa et al. (1984) obtained anisotropic magnets with the composition
Nd 15
15 Fen B
Bea by using a sintering process, while Croat et al. (1984),
Hadjipanayis et al. (1984) and Koon and Das (1984) developed isotropic
permanent magnets from melt-quenched alloys. The basic component of these
alloys is Nd2 Fe14 B, which is tetragonal and exhibits easy-axis anisotropy.
(Herbst et al., 1984; Givord et al.,
ai., 1984; Shoemaker et al. ,1984; Boller
and bsterreicher, 1984).

6.3.3. 1 Phase Diagrams and Crystal Structures


The projection of the liquidus surface in Fig. 6. 30 summarizes all results
al. , 1994). The region of primary solidification of (j) becomes very
(Knoch et ai.,
narrow at higher Nd contents. Nearly all samples show three reactions
between approximately 710 and 670 'c including a metastable transformation.
Existing data of the invariant reactions are listed in Table 6.6.
Nanostructured Exchange-Coupled Magnets 221

40r--------~r\
40~ ~"

Figure 6. 30 Liquidus projection of the system Nd-Fe-8 (Knoch


CKnoch et al. , 1994). Regions of
primary solidification are identified by an underline of the respective phase name.

Table 6.6 Invariant equilibria in the ternary system Nd-Fe-8.


Symbol Reaction T ('C)

PI
PI IS-Fe --y-Fe
L + o-Fe 1392
P,2
P L + y-Fe -- If'(Nd2Fell)
--'/"CNd,Fe'7) 1185
P33 '/"-- NdsFe"
L + If'-- Nds Fel7 780
P, ~-Nd --cx-Nd
L + I3-Nd --ex-Nd 856
Ps IS-Fe -- y-Fe
L + o-Fe 1381
P6a L + Fe8 -- Fe28
Fe,8 1389
P,
P7 L + y-Fe --<I>(Nd Fel,8)
--<DCNd,2Fe" 8) 1180
Paa L + Nd8, -- '1T) (Ndl.l
CNdl.l Fe, 8,) 1345
e, L -- NdsFell
Nds Fe'7 + Nd 685
e,
e2 L -- y-Fe +Fe28
+ Fe, 8 1177
e3 L -- Nd + Nd,8
Nd28ss 1000
e, L -- '1T) (Ndl.l
CNdl.l Fe, 8,) + <I>
(]) (Nd Fel4 8)
CNd,2Fel' 1115
es L --Fe,
--Fe288 + (])CNd,Fel,8)
<I>(Nd2Fe ,4 8) 1110
U,
UI L + Fe8 -- Fe28
Fe, 8 + Nd84
Nd8,
U,2
U L+ Nd8,-- '1T) (Ndl.l
L + Nd84-- CNd"Fe,8,)
Fe4 8,) +Fe,8
+ Fe2 8
U3 L + y-Fe -- <I>
(]) (Nd 2Fel4 8) + If'
CNd,Fel,8) '/" (Nd2Fell)
CNd,Fe'7) 1130
U, Nd,28 ss-- Nd + p (Fe-poor
L + Nd CFe-poor Nd-rich boride)
Us L + p-- Nd + '1T) (Ndl.l
CNdl.l Fe4
Fe, 8 4)
8,) (45)
U6a L + '1T) (Nd1.,Fe
CNdl.l Fe,48 4) -- Nd + (])
8,) <I> (Nd Fe,,8)
CNd,2Fe" 8) 710
E,
EI L -- <I>
(]) (Nd
CNd,2Fe"
Fel' 8) + y-Fe + Fe,
Fe2 8 1105
E,
E2 -- T) (Ndl.lFe
L --'1 CNdl.l Fe,48 4) + <I>
8,) (]) (Nd
CNd,2Fe,,8)
Fel' 8) + Fe,8
Fe2 8 1095
222 W. Liu et al.

According to Schneider et al. (1987), so-called" two-phase" magnets,


which practically belong to the regime of the nominally single-phase magnets
due to including a non-magnetic phase, means that for a given composition
only two phases (L + Nd2Fe148)
Nd2 Fe14 B) coexist at sintering temperature. Since the
shape of the region of primary Nd22Fe14 B 8 crystall ization has changed
drastically, it is necessary to construct a new isothermal section shown in Fig.
6.31 including the situation at 1080 'c, 1050
1080C, 1050C'c and 1000
1000C'c (Knoch et al. ,
1994). The area of primary Nd22 Fe14 B 8 solidification at 1000
1000C'c is extremely
narrow and does not allow us to produce two-phase magnet. However, for the
composition Nd 15 Fen B8 8a ,, one can obtain a two-phase magnet when sintered
above approximately 1080 C 'c ..
FeB
NdlSFenBa
NdlSFenBs
'"
Jo. Nd 18 Fe 7s B6 5
lS ssFe7sB6

/
/ y.
Yo
&~
<llc;
~("o
<;>,..

/
'0
~

/
/
1000'C
60 40 20
Nd (at.%)

Figure 6. 31
6.31 Partial isothermal sections illustrating the area of primary solidification of
Nd 2.Fel4 B (<1J)
Nd,Fel4 (<P) for different temperatures (Knoch et al., 1994).

The Nd22 Fe14 B


8 unit cell is shown in Fig. 6. 32. The lattice symmetry is
P4 2/mnm), and each unit cell contains four formula
tetragonal (space group P4dmnm),
units or 68 atoms. There are six crystallographically distinct iron sites, two
different rare-earth positions and one boron site. In Fig. 6. 32 it can be seen
that each Nd2 2 Fe 14 B
Fe14 8 unit cell consists of an eight-layer repeat structure
perpendicular to the c axis. The R 2 Fe14 B structure has been found to form with
R2Fe148
yttrium, thorium and all the rare-earth elements except europium and
radioactive promethium. For the Nd22Fe14 B 8 compound, the saturation
polarization J s ' the anisotropy field 110 Ha , and the Curie temperature T
/.10 H Toc are
1.60 T, 7. 3 T and 585 K, respectively (Herbst, 1991). The sintered ternary
Nd-Fe-8 magnets with and without additions have high remanence and
Nd-Fe-B
maximum magnetic energy product. Kaneko and Ishihaki (1994) reported a
record of maximum energy product is about 434 kJ/m- 3 for laboratory-scale
magnets. At present, the record for small-scale production is 451 kJ/m 3
(Rodewald et al. ,2002).
, 2002). However the Curie temperature of Nd-Fe-B
Nd-Fe-8 magnets
Nanostructured Exchange-Coupled Magnets 223

CDNd feNd g
eFecOFee ()Fejl ctFej2 eFekl ~Fek2l8>Bg
~Fek2l8iBg

Figure 6. 32 Tetragonal unit cell of Nd


6.32 Nd,2 Fe"
Fe,. B, the prototypical structure of the Nd,
Nd2 Fe,.
Fe!. B
compounds. The c/ciaa ratio in the figure is exaggerated to emphasize the puckering of the
hexagonal iron nets (Herbst, 1991).

is only about 310 C'c ,, which results in large negative temperature coefficients
of the magnetization and the coercivity. This restricts the applications of such
magnets to temperatures below about 50'C. 50C. Therefore, many efforts have
been devoted to overcome the mentioned handicaps. Based on the outstanding
magnetic properties of Nd-Fe-B magnets, efforts have been made to improve
the magnetic properties at room temperature as well as the stability with
respect to reversal external fields and elevated temperatures (Buzro, 1998).
The technical magnetic characteristics including the remanence !Jo /Jo M r , the
intrinsic coercivity /JOM He' and the maximum magnetic product (BH ) max are
!JOM He'

extrinsic and depend crucially on the microstructure of the materials. The


microstructure involving the size, shape and orientation of the crystallites of
the compound and also the nature and distribution of secondary phases, which
usually control domain-wall formation and motion, determines the polarization
and demagnetization behavior. Practical magnets of Nd-Fe-B can be
fabricated by various methods such as conventional powder-metallurgy
processing (sintering) (Sagawa et a!., aI., 1984), melt-spinning (Croat et a!.al. ,
1984), reduction-diffusion (RID) processing (Herget, 1985), mechanical
224 W. Liu et al.

alloying (Schultz et al., ai., 1987) , hydrogen-decrepitation-desorption-


recombination (HDDR) (L' Heritier et al., 1984) and magnetron sputtering
(Cadieu et ai., 1986a, 1986b). To improve the magnetic properties of Nd2 z
Fe14 B-type magnets, much work concerning elemental additives and multiple-
element substitutions has been done. Sagawa and Hirosawa (1987) achieved
(BH ) max ;?: 159 kJ/m 3 and !JOM
(BH)max~159kJ/m3 J.loMH = 5.0 T for the magnet with composition
Hec =5.OT
(Nd oo 53 Dyo
DyO 47)0 15FeO,nBO,08'
15FeO.77B008 Combined Dy and Co substitution is effective in
raising T ec and M M He' and decreasing temperature coefficient ex a and the
irreversible losses (Li et ai.,al., 1986; Tokunaga et al., 1986; Arai et al. ,
1987; Gauder et al., 1988). The highest coercivity obtained in mechanically
alloyed DY18 Fe74 C8 alloy annealed at 900C for 35 min is 9.3 T (Sui et al. ,
1997), although the antiferromagnetic Dy sublattice leads to a strong
magnetization reduction. These experimental results support that Nd2 z Fe14
B-type hard phases may be suitable candidates for the hard magnetic
component in nanocomposite magnets.
6.3.3.2 Exchange-Spring Behavior in Nd-Fe-B-Based Nanocomposite
Magnets
Curie temperatures were measured on a series of melt-spun nanocomposite
materials comprised of Nd2 z Fel4
Fe14 B phase and varying amounts of excess iron
ranging from about 0 to 27 wt. %. The Curie temperatures of the 2 : 14 : 1
phase increase with increasing excess iron; thermal measurements indicate a
13 K enhancement over that of pure Nd2 z Fel4
Fe14 B for 27 wt. % a-Fe,
ex-Fe, while the
magnetic measurements indicate an 18 K enhancement for the same
composition (Lewis et ai.,
al. , 1997).
The average grain size and Curie temperatures for the ribbons of
compositions R9 Fe85 B6 , R117611 76 Fe82,46
Fe8Z46 B5,88 and R18
588 l8 Fe76 B6 (R == Pr and Nd)
prepared by the melt spinning technique are listed in Table 6.7. The changes
of the average grain size are achieved by means of two-step heat treatment.
An enhancement of the Curie temperature for the R2 z Fe14 B (R == Pr and Nd)
phase has been found for exchange-coupled nanocrystalline magnets with
respect to corresponding microcrystalline magnets.
The spin reorientation transition temperature, T sr ' 110 K, of the
Nd2zFe14 B nanophase in the nanocomposite Nd6Fe88 B6 alloy is remarkably lower
NdzFe14B, 135 K (Fang et al. , 1998). Notably, low
than that of single crystal Nd2FeI4B,
temperature spin reorientation originates from the strong interplay between the
crystal-field interactions and exchange interactions on Nd3+ ions. The easy
magnetization direction changes from c axis toward [110] below 135 K for
Nd2zFel4
Fe14 B. Kou et al. (1997) had reported a similar r,esult result that the T sr of a
nanostructured melt-spun Nd2 z Fe14 B single phase alloy is 117 K. The low T sr
might be attributed to the grain size effect of this alloy. The grain size is of the
order of several nanometers and exchange coupling occurs between two
magnetic phases. Furthermore, the temperature for the anisotropy constant
Nanostructured Exchange-Coupled Magnets 225

Table 6.7 The average grain size and Curie temperature of the hard phases measured by
1997>.
aI., 1997).
AC susceptibility for different compositions (Dahlgren et ai.,

Curie
Average Grain Average Grain Curie temperature
Composition temperature
size (nm) size (nm) CC)
COC)
CC)
COC)

Prg Fe85
Fe8S 8 6 32 306 81 295/287(1)
Prl1.
PrlU6 76 Fe82.
Fe82.46
46 8 S.88
5.88 34 309 ?150
~150 294/291(1)
294/291 (I)

Pr18 Fe76 8 6 39 297 ?150


~150 295
NdgFe8s
Fe85 8 6 40 324 70 311
Nd11
Nd ll .76
. 76 Fe82. 46 8 5.88
Fe82.46 S.88 39 321 105 303/318(1)
Nd 18 Fe76 8 6
Nd18 19 298 ?150
~150 318
(1) Two values for Curie temperature correspond to the two peaks in ac-susceptibilty curves.

K11 = 0 may shift to a lower temperature when taking exchange coupling

between the nano-grains into account. Little is known about the origin of the
shift. One possible reason is the limited solubility of Fe in Ndz2 Fe14 B.
Nanoscale exchange effects have also been invoked.
invoked, although there is no hope
to exploit Curie-temperature enhancement effects on length scales of more than
one nanometer (Skomski and Sellmyer.
Sellmyer, 2000).
Wohlfarth (1958) showed that a simple relationship between Md(H) and
M r (H)
M,( H) for non-interacting single domain particles held

Md(H) = Mr(oo) - 2M r (H) (6.20a)

where the remanent magnetization M,(H)


Mr(H) is acquired after the application and
subsequent removal of a direct field H; and the dc demagnetization M dd (H)( H) is

acquired after dc saturation in one direction and the subsequent application and
removal of a direct field H in the reverse direction. Usually.
Usually, the a reduced
form of Eq. (6.20) is adopted,
adopted. in which the remanence polarizations are
normalized with respect to saturation remanent polarization m m,r (00),
(00). i. e.
e .,
=
md(H) = 1 - 2m,(H).
2m r (H). (6.20b)
The Wohlfarth relation (Eqs. 6. 20a or 6. 20b) provides a way to
experimentally monitor the interactions between the grains. For this purpose,
purpose.
usually. the om ( H) plot is constructed,
usually, constructed. (Henkel,
(Henkel. 1964; Kelly et al..
a!., 1989)
which is based on the Wohlfarth relation. om (H) is written
relation, and om(H)
om (H) = m d ( H) - [1 - 2 m rC H) ] (6.21)
(6.21>

where 8mom (H) gives the deviation from the non-intertacting case as a function
of field. This technique has been applied to a wide range of magnetic materials
and used to estimate the type of the dominant coupling mechanism in
particular. thin-film and multilayer samples (O'Grady et a!.
particular, al. , 1993).
om (H)-plots
Several typical 8m (H) -plots are shown in Fig.6.33
Fig. 6.33 (Panagiotopoulos et
al.
al ., 1996). Mainly positive values are present for the single-phase sample.
sample,
226 W. Liu et al.

and this is normally attributed to positive (ferromagnetic) interactions between


the grains. When 8mom values are generally negative, it is usually in situations
where magnetostatic interactions are dominant. In general, however, a
detailed or quantitative interpretation of such plots is difficult.
0.8 0.2

0.1
Nd sFes6 B6
t:
E:
<0
to 0.0
0.0..-1------"..-................ >---
O.a+<----t-::::::-'----
0.0+<-------\0-==-----... --{).I
-0.1

40 50 o0 10 20 30 40 50
H(kOe)
(b)
0.3

0.2 O.OH-------::-=.._......
0.0 >---
t:
E: 0.1
<0
to t:
E:
<0
to -{).2
-0.2
O.
Nd 4 Fe n B I9
--{).I
-0.1
0 10 20 30 40 50 20 30 40 50
H(kOe) H
H (kOe)
(e)
(c) (d)

Figure6.33 8m 5m plots of (a)


Ca) Nd'3 Fe793A107 Si, (b)
NdI3Fe793AI07Si, Cb) Nds FeS6 86, (c)
NdsFes6B6' N~Fes6B6 and
Ce) Nd6FeS686
Nd. Fen B
Nd, 19 (Panagiotopoulos
8 I9 al. , '996).
CPanagiotopoulos et al., 1996).

In real melt-spun composite magnets with non-ideal microstructure, the


coercive field M Heat
He at a given temperature can be described in terms of intrinsic
and microstructural parameters by means of the Kronmuller equation (Bauer
al. , 1996)
et ai.,
(6.22)
where H N is the theoretical nucleation field (= 2K 1/ 1 / M s ) and Ms is the

spontaneous polarization of the hard phase. The parameter ex 0' ~ is related to the
grain orientation (texture) in the magnet; in a randomly oriented Nd2 Fe,. Fe,4 B it is
equal to 0.5. The microstructural parameter ex 0' k takes into account grain
imperfections and Neff M s ((T) T) is an effective demagnetization field acting on
the grains (Neff is effective demagnetization factor). The microstructural
parameter ex = exchange-coupled) is introduced in the nucleation term of
0' ex (ex =
the equation to describe the effect of exchange coupling between neighboring
grains on the coercive field of the magnets. Comparing Eqs. (6. 8) and
(6.22)
(6. 22) reveals that exex -"""- 1/L
1/ L 2 for soft regions of size L .
From the experimental results of the temperature dependence of the
Nanostructured Exchange-Coupled Magnets 227

coercive field the microstructural parameters exKex ex and Neff can be determined
0' KO'ex
by plotting the experimental results M Hel M s vs. the theoretical quantities 0' ex ~
ex
0' K
K ex
0' ex H/M ss Bauer et al. (1996) reported that the values of the
microstructural parameter Neff are significantly smaller for all exchange-coupled
magnets studied as compared to the values for conventional "uncoupled" melt
spun and sintered magnets. The rather small values of O'ex ex ex in the remanence-
enhanced magnets illustrate the drastic influence of random exchange coupling
on the coercive field of these magnets. Therefore, the exchange interaction
between the grains is responsible for the remanence enhancement, and the
exchange hardening of the soft magnetic phase limits the coercive field of
exchange-coupled magnets. Billoni et al. (1998) studied the dependence of
the coercivity on the grain size in a NdFeB : ex-Fe nanocomposite. These
results suggest that in the nanocomposites characteristic of exchange coupl coupling,
ing,
the parameter 0' ex ex
ex depends not only on the volume fraction of soft phase but

also on grain size and grain-size distribution.


Figure 6.34 shows the dependence of coercivity on grain size for different
volume ratio fractions of hard and soft phases of nanocomposite magnets of
NdzFe14 B : ex-Fe (Sun et al. , 1999). The results confirm that the coercivity of
the nanocomposite magnets depends on the grain size and on the volume
fraction of soft phase. In fact, the experimental results show that the addition
of ex-Fe leads to magnetic softening of the nanocomposite magnet magnet..
...... 10:0
.......
-0-9:1
-9:\
--8:2
--- 8:2
2.5 -0-7:3
....... 6:4
~ 2.0 -6-
-l:r- 5:5
b .........
--4:6 4:6
j:1.5 -<:;-
-<:l- 3:7

1.0

0.5~~====-
0.5~~===-
10 15 20 25 30 35 40
a (nm)

Figure 6. 34 Computational dependence of coercivity on grain size for different


'h : Ndz Fe" Sf
composition ratio f h : f'ss of nanocomposite magnets of NdlFe14 'h
ex-Fe, here f hand
B/a-Fe, fs
and's
represent volume fractions of hard and soft phases, respectively (Sun et al. ,1999).
,1999) .

The magnetic hardness of the nanocomposite magnets originates from the


contributions of both the hard magnetic and the soft magnetic components, as
in the case of the rare earth-transition metallic compounds in which the
anisotropy of the compounds comes from the contributions of both the rare-
earth and transition-metal sublattices. In order to characterize the hardness of
the nanocomposite magnets, a descriptive physical quantity, effective
228 W. Liu et al.

anisotropy K eft'
eff' was suggested by Skomski and Coey (1993) and Sun et al.

1999 ), from theoretical and experimental standpoints, respectively. The


( 1999),
concept, which is limited to very small grain sizes, amounts to a resultant
= tf s K s + tf h K h' where tf s ' ft h and K s' K h represent the
eff =
anisotropy constant K eft
volume fraction, anisotropy constant of soft and hard components,
respectively.

6.3.3.3 Nd 2 Fe 14 B: Fe3B Nanocomposite Magnets


Nd2Fe14B:
Coehoorn et al. (1988) first synthesized a novel nanocomposite magnet
containing hard and soft magnetic phases. The material with the nominal
composition Ndu Fen B 18 .5 was prepared by rapid-quenching and annealing of
melt-spun ribbon to yield a mixture containing a Fe3 B-rich phase, Nd2 2 Fe'4
Fel4 B
and a small fraction of ex-Fe. The Fe3 B has a saturation induction of 1.6 T at
room temperature, planar anisotropy and behaves as a soft magnet (Coene
et al. , 1988). The Curie temperature of Fe3 B is 786 K, nearly 200 K higher
than that of Nd2 Fel4 B. The exchange coupling coupl ing between the fine Fe3BFe3 Band and
Nd2Fe'4
Fel4 B grains results in a remanence enhancement.
Coehoorn et al. (1988) reported that after annealing anneal ing amorphous
Nd 4Fe78 B 18 an isotropic permanent magnet with B,
Nd4Fe78BI8 Br == 1. 2 T and (BH)
(BH)max max =
=
86.4 kJ/m 3 3
was obtained. For a similar alloy, Shen et al. (1990) obtained
(BH)max = 106.8 kJ/m 3 . Li et al. (1992) achieved the optimum magnetic
= 106.8kJ/m
properties: B, = 1. 2 T and (BH ) max ~ 100.8 kJ/m 3 in Nd4Fen B 1919 ribbon. The
Br =
common characteristic of the materials is that the crystallized materials are
composites consisting of Nd2 2 Fel4
Fe14 B, Fe3 B and ex-Fe phases in nano-scale, and
exhibit a remarkable remanence enhancement.
The composite of Nd4Fe78 Fe7B B 18 , containing over 50 % Fe3 B, has /Jo Mr ~
J.io M,
1. 2 T, /JOM
1.2T, J.ioMH = 0.38 T and (BH)max~
Hec =0.38T 96 kJ/m 3 (Coehoorn et al.,
(BH ) max ~ 96kJ/m aI., 1989,
Hirosawa et al. , 1993). A small change in composition, namely Nd4Fen5B'85 Nd4Fen5BI85
leads to an energy product of 104 kJ/m 3 (Zhang et al. , 1990). The melt-spun
alloy of Nd2 Fe23 B3 (Nd 71 Fe82. Fe8211 B IO 7) annealed under optimum conditions has
been found to be composed of ex-Fe, Fe3 Band Nd2Fel4 Fe14 B phases. The alloy has
a coercivity of about O. 026 T and relatively high-energy product of about
71.6 kJ/m 3 (Kim et al., aI., 1995). The metastable compound Nd6Fen Nd6Fen B 17 has
been obtained by the crystallization from the amorphous phase in the Nd-Fe-B
system. Its crystal structure is similar to that of the compound Nd2Fe23 B3 . At
high temperature this metastable compound decomposes into ex-Fe and
Nd l . I, Fe4 B4 phases. The room-temperature magnetization and coercivity are
Nd,.
154 Am 2/kg and 0.015 O. 015 T, respectively (Gu et al. , 1991).
Rx Nd4- x Fen 5B 185 , the substitution
Cheng et al. (1995a) reported that in RxNd4-xFen.5BI85'
of R (R = Pr, Gd, Dy, Y) for Nd results in decreasing of the coercivity with
(R=Pr,
increasing rare earth (R) concentration for R = = Pr, Gd, Y. The addition of a
small amount of Dy will improve the coercivity but full substitution of Dy will
result in a decrease of the coercivity owing to the presence of the magnetically
soft phase R3Fe62 B 14 . The addition of Gd and Dy causes an increase in the II11 B
Nanostructured Exchange-Coupled Magnets 229

hyperfine field in Fe3 B but does not influence that of the ex-Fe. This can be
explained by the entry of R atoms into Fe3 B. They also investigated the effect
of Nd content on hard magnetic properties of (Fe3 B) B),_1- x (Nd2 z Fe14 B) x alloys
(x=0.17, 0.26, 0.34, 0.42,0.65,0.78
0.42, 0.65, 0.78 and 1.0) by rapid quenching and
post heat treatment (Cheng et al., 1995b). After an appropriate heat
treatment, melt spun Nd4Fen5B18.5
Nd4FenSBI8.S alloy has a coercive field of 0.30 T, a
remanence of 1. 25 T and an energy product of 104 kJI m3. The results indicate
that the samples with higher Nd concentration consist of Nd2 zFe14 B,
Nd l l Fe4 B4, and a small amount of ex-Fe. The melt spun Nd-Fe-B alloys with a
low Nd concentration annealed under an optimal heat treatment consist of
body-centered-tetragonal Fe3B Fe3 B (bct-Fe3
(bct-Fe3B)B) and a few percent of ex- Fe and no
Nd2Fe14B magnetically hard phase; about 5 at. % Fe atoms in Fe3 (8g) site of
NdzFe14B
bct-Fe3 B are replaced by Nd atoms in the samples annealed under optimal
conditions. Results of nuclear magnetic resonance (NMR) radio frequency
enhancement effect demonstrate that Nd-containing bct-Fe3 B has better
permanent magnetic properties. The m6ssbauer effect results on the
substitution of Pr, Gd and Dy for Nd confirmed that the remanence
enhancement of Nd Nd4Fen5B185
4Fen SB 18 S alloy annealed at 670C for 2 min originates
from the strong exchange coupling
coupl ing between the Nd-containing bct-Fe3 B with
uniaxial magnetocrystalline anisotropy and soft magnetic ex-Fe with a nanoscale
(about 30 nm). The hard magnetic properties are due to the existence of Nd-
containing bct-Fe3 B. They found that two types of Fe3 B phases including
o-Fe3 Band
O-Fe3B and bct-Fe3
bct-Fe3B,B, are formed in Nd4Fe77sB,8s-xCx(x
Nd4Fen.5 B 185 - x C x (x =0,4,6,8
= 0, 4, 6, 8 and
10) alloys. The substitution of carbon for boron atoms favors the formation of
o-Fe3 B phase. The decrease of coercivity is attributed to the replacement of
O-Fe3
the bct-Fe3B
bct-Fe3 B phase by O-Fe3
o-Fe3B ones (Cheng et al. , 1995c).
Bones
Zhao et al. (1999) and Xiao et al. (2001 a) reported that partial
substitution of Sm
8m for Nd could be favorable for the exchange coupling between
soft and hard phases due to anisotropy reduction in alloy ribbons with the
Nd4- x Sm x Fen 5B 18 5(x = 0.0
composition Nd4-x8mxFensBl8S O. 0 - 0.5).
O. 5). According to the exchange-
spring model of Kneller and Hawig (1991>,(1991), the correlation length for the soft-
magnetic phase can be expressed as bern = "IT (A 12K) 1/2,
b em = I/Z, where A is the
exchange energy of the soft-magnetic phase, and K is the magnetocrystalline-
magnetocrystall ine-
anisotropy constant of the hard-magnetic phase. To obtain a sufficiently strong
exchange coupling, the grain size 0 of the soft-magnetic phase must be
smaller than 2b cm em . However, this approach reduces the coercivity without

enhancing the remanence; it cannot be used for permanent magnets.


Kanekiyo et al. (1993) reported that M = AI, Si, 8i, Ga, Ag and Au
additives reduce the grain size and contribute to the improved magnetic
performance of rapidly solidified, crystallized, and resin-bonded
NdsFe70S Cos B I85 M, magnets where M is any of the above-mentioned
Nd5Fe70.5C05BI8.5Ml
additives. In melt-spun Nd5 s Fe74 Cr3 B 18 , Cr induces a desired scheme of
decomposing the intermediate compound Nd2 z Fe23
FeZ3 B3 into Nd2 z Fe14 B, Fe3 Band
ex-Fe, whereas for the corresponding Cr-free alloy Nds5Fen B 18 it decomposes
230 W. Liu et al.

along with Fe3 B into another mixture composed of a- Fe and non-magnetic


NdFe
NdFe44B4 without any hard magnetic phases <Uehara et al., aI., 1998a, 1998b).
Thus, the substitution of Cr for Fe allows the existence of a large fraction of the
hard-magnetic Ndz2 Fel4 Fe14 B phase in the nanocomposite structure, resulting in
increasing coercivity.
Gao et al. (1998a) investigated the nucleation and magnetic interaction in
multi phase nanocrystalline Nd4 Fen C03B I6 and Nd4 Fe76 C03 (Hf 1 l - x Ga x ) B I6
(x = 0, 0.5 and 1). The results showed that a nucleation process controls the
coercivity of the magnet. The oM 8M plots show that both exchange-coupled
interaction and magnetostatic interaction in the sample N~ F~6 C03Hf05 Ga055B
o 5Gao B16 are
stronger than those in Nd4Fen C03B 16 . They found that the magnetic field heat
treatment not only induces grain refinement but also causes a uniform
distribution of the soft and hard phases. The existence of amorphous phase
reduces the exchange-coupled interaction between the two magnetic phases
and results in a relatively low coercivity in the nanocrystalline magnet. The
exchange-coupling of the sample Nd4 Fe76 C03Hfo .5 Gao
o5 GaO.55 B I6 was greatly
enhanced in the sample annealed with magnetic heat treatment, achieving the
highest energy product (BH)max = 126 kJ/m 3.
In melt-spun Ndu Fen Fe73 C03Gal
Ga, B 18 5 alloy, Co and Ga atoms are rejected
I8 .5
from primary particles of the soft magnetic Fe3 B phase and partitioned to the
amorphous matrix enriching at the Fe3 B/amorphous interfaces in the early
stage of crystallization; three crystalline phases (Fe3 B, Nd2
(Fe3B, FeZ3 B3 and
z Fe23
NdzFe14B) are present in the fully crystallized state. For the fully crystallized
Nd2FeI4B)
Fe3 B : Nd2 zFel4 B nanocomposite with optimum magnetic hardness, Co and Ga
atoms are partitioned to the Nd2 z Fe
Fe14 14 B phase, and evidence for a slight
enrichment of Ga atoms at the Fe3 B : Nd2 z Fe14 B interface was found (Ping
et al. , 1998). They also studied the effect of Cu and Nb on the microstructure
and magnetic properties of an Fe3 B/Ndz2Fel4 B nanocomposite magnet (Ping
et al. , 1999). Trace addition of Cu (0.2 (0. 2 at. %) is very effective in reducing
the grain size of the nanocomposite microstructure. A combined addition of Cu
and Nb further reduces the grain size. The optimum magnetic properties of
/Jo M r = 1.25 T, /JOM
JJo M, JJoMHHec = 0.34 T and (BH)max = 125 kJ/m 3 were obtained in a
NduFe758BIssCuozNbi
NduFe758BI85CUQ2Nbl alloy annealed at 660 660C'c for 6 min.
Yang et al. (2000) investigated the effect of Co in Nd4Fe775-x-zyCox(Hf,
Nd4Fe775-x-2yCox(Hf,
Ga) yB 185
Ga)yB 18 5(0 :s;; x ::(
(0::( :s;; 5, yY = 0, 0.5)
O. 5) on the magnetic properties of Fe3B Fe3 B :
Ndz2Fe14
Fel4 B magnets. It was found that Co retarded the crystall ization of (X-Fe a-Fe or
Fe3 B but accelerated that of Nd2 zFe14
Fel4 B. Thus, the interval between the onset of
crystallization of Fe3 Band Nd2 z Fe14
Fel4 B phases is decreased, enabling the grain
growth of each phase to be uniform during a post-annealing of the melt-spun
ribbons. From the ribbon magnets of Nd4 Fe71 5 C05 Hf05 GaO 5 B I85 made at
B'85
26 m/s and annealed at 680C 680 'c for 10 min, the magnetic properties of /Jo M,r =
JJoM
1.15T, /JoMHJJoMH ec =0.35Tand (BH)max=114kJ/m 3 3
were obtained.
Bernardi et al. (1999) considered (Nd,Tb)55(Fe,Cr,M)765B18(M=Co,
(Nd, Tb) 5 5(Fe, Cr, M) 765 B 18 (M = Co,
Si) nanocomposite prepared by splat cooling and subsequent fast annealing
Nanostructured Exchange-Coupled Magnets 231

between 700 and 800 C. They found that optimum magnetic properties were
achieved for the sample containing 3 at. % Cr. A higher Cr content results in
an increase of the coercivity but also in a decrease of both remanence and
saturation magnetization. The magnetic properties of the splat-cooled
materials with 3 at. % Cr can be improved by the addition of Co or (Co, Si).
After optimized annealing the magnetic materials consist of a homogeneous
microstructure of Fe3B and (Nd,Tb)2Fe14B grains with a diameter of about 10
- 30 nm. The splat cooled flakes with 0.5 at. % Si and 2.5 2. 5 at. % Co show a
remanence of 1J0 M, =0.
IJoM, = 0.91
91 T and a coercivity of IJoMH IJOM Hec = 0.69 T.

Mishra and Panchanathan ((1994) 1994) investigated the microstructure and


magnetic properties of bonded and fully dense magnets produced from melt-
spun ribbons of the composition Ru4 5T76Gal T 76 Gal B 185
18 5', where R refers to a mixture
of Nd and Dy, and T refers to a mixture of Fe and Co. The results show that
annealing overquenched ribbons with 3 wt. % Co at 700 C can produce
materials with remanence 1J0 IJ oMr""'" 1 . 15 T, intrinsic coercivity 1J0
M r "'1. M He""'" 0 . 4 T
IJoMHc"'O.
and a maximum magnetic energy product of about 116 kJ/m 3 , being
comparable to those of commercial ribbons containing three times as much
Nd. The microstructure of the annealed ribbons consists of about 30 vol. %
Nd2Fe14B grains, 65 vol. % Fe3B and 5 vol. % a-Fe ex-Fe grains. The grains of
all the phases are nearly spherical and have diameters in the 30 - 50 nm
range.
Zhang et al. (zhang and Inoue 2000; Zhang et al., 2000b, 2001 b )
reported a glass transition in melt-spun Ndy DY05 Fe995-x- y-zz Cox B z (x = 4 -
Fe995- x- y-
56, y = 4 - 4. 5, zZ = 18 - 30) amorphous alloys. The crystall ized structure
consists of Fe3 B, Nd2 Fe14 B, a-Fe ex-Fe and remaining amorphous phases in
annealing temperature of up to 930 K for 420 s and changes to Fe3 B Nd2Fe14 B
ex-Fe for higher temperatures. The grain size after annealing at 903 K is
and a-Fe
about 20 nm for Fe3 B, 10 nm for Nd2 Fe14 B, 30 nm for a-Fe ex-Fe and the
interparticle spacing among their crystalline phases occupied by the remaining
amorphous phase is about 5 nm. The polarization at a field of 1.6 T,
remanence, intrinsic coercivity and (BH) max are 1. 36 T, O. 29 T and 110 kJ/
m3 , respectively, for the amorphous Nd3DY05 Fe67 Cogs B20 alloy annealed at
903 K for 420 s. The good hard magnetic properties result from the exchange
coupling among the four ferromagnetic phases of Nd2Fe14 B, Fe3 B, a-Fe ex-Fe and
remaining amorphous phase at room temperature. The magnetic properties of
Nd2Fe14 B/Fe3 B type nanocomposite magnets with different compositions and
synthesis are listed in Table 6.8.
It is concluded that the coercivity and energy product are relatively low
due to the relatively low saturation magnetization of soft phase Fe3 B in
Nd2Fe14B:
Nd 2Fe14 B : Fe3B
Fe3 B type nanocomposite magnets.
6.3.3.4 Nd 2 Fe I4
14 B: a-Fe Nanocomposite Magnets

Since the exchange-coupled magnets were reported, it has been found that as
a soft phase, Fe3 B is not unique in two-phase nanocomposite magnets, since
232 W. Liu et al.

its saturation induction of 1.6 T is in fact lower than that of ex-Fe, and ex-Fe is
considered as the soft phase in most of investigations on exchange-coupled
magnets.

Table 6.8 Remanence, /.laM" M,IM s and maximum


/JoM" coercivity, /JoMH cc '' remanence ratio, M,/M
energy product C max of Nd
BH )max
(BH) Nd,2 Fe" B/Fe3 B type nanocompsite magnets with different
compositions, and with different synthesis and with different 3d element additions.
/Jo M,
/.laM, /JoMH c M, CBH)max
Compositions Synthesis Ref.
CT)
(T) (T) IM s
/M CkJ/m 3)
Nd, Fe80 B 16
'6 Rapid quenching -1.28 -031 0.84 115 [IJ
C03Hfo. 5GaO.5
Nd, Fe76 C03Hf05 Gao 5BB'6
16 Rapid quenching -121 -0.29 0.84 122 [2J
Nd,
Nd4.55Fe75.
Fe75.88 B 18
18 .. 5CUo.
5Cuo,2Nb,
Nb 1 Rapid quenching 1.25 0.34 125 [3J
Nd,Fe 715 C0
Nd, Fe7l5 5HfQ.5Gao
C05Hfo.s 5B 18.5
Gao 5B'8.5 Rapid quenching 1.15
115 036
0.36 114 [4J
Nd, Fen5
Fe735 C0 3Hlo.5Gao 5B 185
Hf05 I85 Rapid quenching 1.25 0.28 -0.81
028 126 [5J
C03Gal B'8.5
Nd4.5 Fe73 C03Ga, B 185 Rapid quenching 1.20 0.43
043 128 [6J
Nd3Dyo COg955B,o
Dyo. 5Fe67 C0 B20 Rapid quenching 1.36
136 029
0.29 110 [7]
Nd3.5Dy,
35 DYI Fe73 C03Ga,
Gal B I85
185 Rapid quenching 115
1.15 050 132
[lJ aJ., 1998b; [2J Ping et aI.,
[IJ Gao et aI., aJ., 1999; [3J Yang et aI.,
aJ., 2000; [4J Yang and Park.
Park, 1997;
[5J Kanekiyo et aI.,
aJ. .1993; al. , 2001b; [7] Kanekiyo and Hirosawa, 1998.
1993; [6J Zhang et aJ.

Davies et al. (1993a) studied the magnetic property of nanoscale


NdxFe94-xB6 8 ~ x ::::;;;
Nd x Fe9'- x B6 alloys with 8::::;;; ~ 12. As x decreases below about 11 at. % %,,
ex-Fe as second-phase appears and the volume fraction of ex-Fe grows up to
about 35 vol. % at x = = 8 at. %. This result is a further enhancement of the
remanence compared to nanoscale Nd2 Fe ,4 B (up to about 1. 1 T at x = 8)
Fe,4
because of exchange coupling between the Nd2Fe,4BNd2Fe14B and ex-Fe grains. The ex-
Fe grains are about half the size of the Nd2 Fe,4 Fe14 B, typically 15 nm, and the
ribbons remain isotropic. As explained in Section 6. 2, because the ex-Fe is
present on an ultra fine scale, it does not result in serious deterioration of the
loop shape in the second quadrant. In fact, (BH) max is enhanced up to about
160 kJ/
kJ/mm3 . The work of Davies et al. (1993b) confirmed that, by careful
control of the melt-spinning conditions, refinement of the grain size in melt-
spun Nd-Fe-B alloy ribbon containing a small concentration (1. 2 %) of sil icon
results in enhancement of remanence and maximum magnetic energy product
and that the effect commences at a mean threshold grain size of about 35 nm.
The smaller the grain diameter is, the greater the improvement in remanence.
The increase in (BH)max appears to be limited to approximately 160 kJ/m 3 for
individual Nd-Fe-B ribbons, due to the attenuation of coercivity that also
results from exchange coupl ing.
Manaf et al. (1993a, b) reported that the best results, J./o M,r = 1.
110 M 1. 13 T,
J./oMH c =
110MH = --0.61
0.61 T, and (BH)max = = 163 kJ/m3 , are achieved in alloys with 8-
10 at. % Nd and 5 - 6 at. % B. With respect to microstructure, the alloys
consist of a matrix of magnetically hard Nd2 Fe14 B phase with numerous
zFe,4
particles of ex-Fe at the grain boundaries or vice versa, depending on the
Nanostructured Exchange-Coupled Magnets 233

relative content of soft phases. Both the mean grain size of the the' Nd2Fe14 B
(smaller than 30 nm) and the mean particle size of a-Fe ex-Fe (smaller than 10 nm)
are ultrafine.
ultrafine"
Withanawasam et al. (1994) reported that a reduced remanence as high
0" 78 with saturation magnetization 190 Am 2/kg led to (BH)
as 0.78 (BH)max kJI m3
max of 111 kJ/m
3

in Nd2Fe14 B : a-Feex-Fe nanocomposite prepared by melt-spinning, with more than


wt" % a-Fe.
50 wt. ex-Fe. The nanostructure composed of a-Fe, ex-Fe, Nd2Fe14 B and remaining
amorphous phases is formed for Fe-rich Nd-Fe-B amorphous alloys annealed at
923 - 1023 K for 60 - 300 s. Rather good hard magnetic properties, i. e.
(J.loM,) of 1. 28 T, coercivity of 0"
remanence (110M,) O. 32 T t and maximum energy
product (BH)max of 146 kJ/m for Nd7 Fes9 3
3
Fe89 B4 are obtained. The Nd2 Fe14 B
particles with a size of about 30 nm are surrounded by the a-Fe ex-Fe and amorphous
phases that act as a magnetic exchange-coupled medium" medium. The nanoscale
coexistence of three ferromagnetic phases is important for the achievement of
the rather good magnetic properties in this system (Inoue et al. , 1995).
Chang and Hsing. (1996) reported that the homogeneous grain
morphologies were achieved in the Nd 12 - x Fes2+ =
Fe82+ x B6 for x = 0 prepared by rapid
quenching at the lower wheel speed Sw (18 m/s :::;; :::;;; Sw:::;;
Sw:::;;; 25 m/s), which
cause a strong exchange coupl couplinging effect between Nd2Fe14 B and a-Fe ex-Fe grains. In
the alloys with x = 2.5, 3 and 4, exchange coupling is also found and the best
magnetic properties 110M, 1. 14 T, 110M
J.loM r = 1.14 J.lOM He = 0.63 T and (BH)max =
kJI m3
151 kJ/m 3
are achieved in the alloy ribbons with x = 2. 2.55 quenched at a wheel
speed of 18 m/s ml s without any subsequent heat treatment. treatment"
Starting from nearly single-phase Nd2Fe14 B magnets, the contents of ex-Fe a-Fe
in composite magnets were stepwise increased up to 40 vol. % by reducing the
Nd and B content. The maximum remanence of 110 J.lo M =
M,r = 1.25 T, the coercivity
110 M He =
J.lOM = 0. 53 T and the maximum energy product (BH)
0.53 (BH)max max = = 185.2 kJI m3
185. 2 kJ/m 3
were
achieved in nanocomposite magnets containing 30 vol. % a-Fe ex-Fe (as shown in
'dimensions of about 4 mm x 3 mm x 0.03 mm3
Fig. 6.35). (The flakes had "dimensions 3
and
the demagnetization factors N was calculated using an expression valid for
ellipsoids. The values of N were on the order of O. 01 ). A representative
micrograph of a composite Nd 10 lO Fes4 (14.22 vol. % a-Fe)
Fe84 B6 magnet (14" ex-Fe) is shown in
6" 36. The microstructure investigations of the composite magnets reveal
Fig. 6.36.
two characteristic maxima in the grain distribution at about 15 nm and 25 nm
corresponding to the a-Fe ex-Fe phase and to the Nd2 Fe14 B phase, respectively. respectively"
The high-resolution TEM micrograph of a composite Nd ,o lO Fe84
Fes4 B 6 magnet is shown
in Fig. 6. 37. A typical grain arrangement in two-phase magnets clearly proves that
the small grains represent the ex-Fea-Fe phase and the large ones represent N~ Fe14 B. In
the bigger N~ Fel4Fe14 B grain the marked crystal spacings denote the basal plane of the
tetragonal unit cell. The smaller a-Fe ex-Fe grain partly covered by the N~ Fe'4 Fe14 B grain
shows in the darker part the (110) planes with the characteristic lattice spacing of
2.02 A. (Bauer et al., aI., 1996).
Starting from rapidly quenched Nd 6 Fe13.1 B (2.05
Nd6Fe131B (2.05:::;;; :::;; 5:::;; 147" 6) alloys
5:::;;; 147.6)
and by appropriate annealing,
anneal ing, the microstructure was tailored from strongly
234 W. Liu et al.

2 200
0.... .. 0 .......
00 .... ..oo.o '6. ~BH),:
..0.. 00 ... ....(BH)--
..... o'b.. max
0 ..
o. --.
~

E ~

~
~
..... -I- --1-" "'t-" - .;ljl
u - - -+- --JR 100 6 x
::t;0
:r:: "
~E
~

:: ~
:r::
--}J.oHc
~

0
0 10 20 30 40
Content of a-Fe (vol.%)

Figure 6. 35 Dependence of magnetic properties of NdFeB-based nanocomposite magnets


on a-Fe concentration at room temperature (Bauer et al. , 1996).

Figure 6. 36 TEM micrograph of melt - spun Nd ,OlO Fe"


Fes. B6 ribbon flake (14 2 vol. % a-Fe)
(Bauer et ai., 1996).

interacting Nd2 2 Fe'4


Fe!4 B grains to magnetically isolated Nd2
2 Fe!.
Fe'4 B grains in a Nd-
rich matrix (ex-Nd
Cex-Nd and y-Nd). This change in microstructure was found to have
a large effect on coercivity, i. e. , coercivity increases with an increase of the
Nd concentration from 1. 25 T in Nd205 Fel31
Fel3.! B to 2.75 T in NdNd,47
'47 6Fel3,
6 Fel31 B at

295 K. It is concluded that the interaction between the grains of Nd2 2 Fe'4
Fe,4 B
phase leads to the decrease of the coercivity in this system. The largest
observed coercivity, about 2.75 T in such nanocomposite Nd Nd,47.6
,47 .6 Fel3.,
Fe,3.1 B
alloys, is about 83 83%% of the theoretical limit expected for Stoner-Wohlfarth
coherent rotation behavior including demagnetization effects (Girt CGirt et al.,
2000, 2001).
To Summarize, the best result for the maximum energy product is 163 kJ/m3 3
in
Nd2Fe,4B:
Nd2 Fe!4 B : ex-Fe nanocomposite prepared by rapid solidification. Meanwhile,
in order to improve the magnetic properties, much work on Nd2 2 Fe!4
Fe'4 B : ex-Fe
type nanocomposites with additional 3d elements has been done.

6.3.3.5 Nd 2 Fe ,4I4 B: a-Fe Nanocomposites with Addition of 3d Elements


Wecker et al. (1995)
C1995) investigated nanostructured isotropic Nd-Fe-B magnets
prepared by hot compaction of mechanically alloyed Nd-Fe-B powders at
Nanostructured Exchange-Coupled Magnets 235

(a)

(b)

Figure 6. 37 High resolution TEM micrograph of a Nd 10


lO Fe84 B6 ribbon flake (14. 2 vol. %

ex-Fe) (Bauer et ai.,


a-Fe) aI., 1996).

temperature of about 600 C. 'C. The best results are remanence of 1. 0 T,


maximum magnetic energy product of 121 kJ/m 3 and coercivity of 0.53 T for
Nd-Fe-Co-Si-B magnet.
Yao et al. (1994) studied coercivity of Ti-modified nanocrystalline
(ex-Fe) -Nd 2Fe14 B alloys with iron particles embedded in a hard Nd2 Fe
(a-Fe)-Nd2Fe14B 14 B
Fe14
matrix, prepared by melt spinning and annealing. The grain size of Nd2 2 Fe14 B is
from 20 nm to 60 nm, while that of the soft iron particles is 11 nm to 30 nm.
For the Ti-alloyed flakes, the 1 at. % Ti-containing ones show the highest
coercivityof 1. 1 T. Its (BH ))max max value is about 127. 4 kJ/m 3 despite
substoichiometric Nd and B contents. More Ti addition would deteriorate the
coercivity. The addition of Ti is favorable for enhancement of the intrinsic
coercivity in mechanically alloyed Nd-Fe-B alloys with a low level of Nd.
Specifically, the coercivity of Nd lO
10 Fe76 B2Ti 12 with nearly single Nd2Fe14 B type
phase is over O. 98 T, and the increase of the Curie temperature in the Ti-
doped alloys with an excessively low B content imply that part of the Ti atoms
may not occupy crystalline sites of the Fe and Nd sites (Zhang et ai., aI., 1998,
Liu et al. , 2000c).
236 W. Liu et al.

Exchange coupling between magnetically hard and soft phases in annealed


melt-spun Nd 6Fes8- x M xB6 alloys (M =
Nd6Fess-xMxB6 Ti or V, x =0-5)
=Ti = 0 - 5) was studied by Fang
and Chin (1996). Addition of Ti or V to replace iron is effective in increasing
the reduced remanence, M,j M,/M M s ' and the energy product of the modified
alloys. Studies of high-resolution TEM showed that both Ti or V addition are
effective in refining the grain size of both hard and soft phases to 20 - 23 nm
compared with 50 - 52 nm for x = 0 alloy. To achieve a higher coercivity, Ti
addition is a good choice; for both a higher remanence ratio and (BH) max' V
addition is preferred.
Withanawasam et al. (1996a) reported that Nd x Fe93-x Fe93- x Nb ll B6 melt-spun
ribbons with x ~ :?o 10 crystall ize directly from amorphous to a mixture of
crystallize
Nd2Fe14B and ex-Fe phases. For 7 < x < 10, the final mixture of 2 : 14 : 1
Nd2Fel4B
and ex-Fe phases is obtained through a two-stage crystallization where the
intermediate metastable Nd2Fe23 B3 and Nd3Fe62 B I4 14 phases are produced at the
initial crystall ization. In samples with 4 < x < 7, the intermediate phase is
crystallization.
Nd3Fe62 B 14 No metastable phases were observed in samples containing less
than 3 at. % Nd although the calorimetric measurements indicated a two-step
crystallization process.
Hadj ipanayis et al. (1995) reported the crystall ization, crystal structure
Hadjipanayis
and magnetic properties of R6Fes7 Nb)l B6 (R = Nd, Pr, Dy, Tb) and
FeS? Nb
Nd3 .5 Fe91 Nb2B3 5 melt-spun ribbons with a microstructure consisting of a
35
mixture of exchange-coupled magnetically hard R2 Fel4 Fe14 B and soft ex-Fe type
phases. At first the Y3 Fe62 B I4 14 -type + ex-Fe phases are formed for R = Nd, Pr
and Dy and subsequently they transform to 2 : 14 : 1 and ex-Fe upon heating
700 'C. The intermediate phase in the case of Tb 6Fes7 FeS? Nb l B6 is of the TbCu7-
TbCu?-
type. A high remanence for Nd3 355 Fe91 Nb 2B3 5 ribbon with a reduced remanence
of M,/ M s of up to 0.8 was observed. A room temperature coercivity of 0.45 T
was obtained in the Nd4Tb 2Fe86 Fes6 Nb 2B6 sample.
In order to improve magnetic performance, Crespo et al. (1997)
investigated the influence of Si, Zr or Cu in NdsFes7B4X NdsFeS? B4X (X=(X = Cu, Zr, Si) and
Nds Fes B
Fes44 4 3Nb Cu. The maximum energy product is achieved in Nd8 Fes8 B4
s Fess
prepared by mechanical alloying and a solid-state reaction at temperatures
around 650 'c and the values are reasonably good for the addition-free
composition and the additions of Zr and Si. The coercive force could be
improved with the additions of Zr and Nb 3Cu, while Si does not yield an
increase compared with NdsFes8 Fess B4. The remanence values for these
compositions are unchanged in the case of Zr, somewhat higher when adding
Si and lower for NdsFes Fes44B4Nb 3Cu. The magnetic behavior can be improved by
introducing a small amount of additives, such as Zr and Si, which avoid a large
increase of the grain size with annealing.
anneal ing.
Miao et al. (1997) showed that the structure and magnetic properties of
mechanically milled Nds (Fel- xCoxx )ssB
(Fel-xCo )ss B4 (x = 0 - 0.6)
O. 6) depend on composition
and heat treatment conditions. Samples with 10% - 20 % substitution of Co for
Fe and heat treatment at 600 'c exhibit a uniform grain size of about 20 nm and
Nanostructured Exchange-Coupled Magnets 237

significant remanence enhancement. A remanence of over 120 Am 2/kg and a


coercivity of about 0.50O. 50 T were obtained in optimally annealed samples.
Excessive substitution of Co for Fe and heat treatment at temperatures higher
than 600'C
600C caused the deterioration of both remanence and coercivity.
The principal advantage of the Co addition rests with an increase in the
Curie temperature of the hard phase and intrinsic coercivity, and an increase
in the saturation polarization arising from the higher saturation of ex-(Fe, a-( Fe, Co)
compared to that of pure ex-Fe.a-Fe. For Nd 126
12 .6 Fe678
Fe67.8 CO I16
ll 6
. Cr2 B 6 : ex-Fe
a-Fe with a
37.5 vol. % a-Fe,ex-Fe, prepared by ball milling and annealing, a remanent
magnetic polarization of 1. 1 T and an intrinsic coercivity of O. 96 T were
achieved by Jurczyk and Jakubowicz (1998).
The influence of Co substitution in nanocrystalline exchange coupled
Nd2(Fel-xCox)14B Nd2 (Fel- x Cox) 14 B + (Fel-
Nd2(Fel- x Cox) 14 B single phase materials, Nd2(Fel-xCox)14B+ x Cox)
(Fel-xCox)
composite magnets and exchange decoupled Nd2(Fel- xx Cox) 14 B magnets with
overstoichiometric Nd (all with x = O. 12 and O. 24) have been studied by
Melsheimer et al. (1999). Both single phase and composite magnets show a
significant remanence enhancement and a high maximum magnetic energy
175. 8 kJ/m 3 , compared to 93.3
products of up to 175.8 93. 3 kJ/m 3 for the exchange-
decoupled magnet (both at 300 K). By substituting Co for Fe in Nd2Fe14 B the
Curie temperature rises to 780 K for the 20 at. % Co-substituted material and
the maximum energy products of almost 120 kJ/ kJ/mm3 are achieved at 435 K. The
improvements in the magnetic properties of as-milled Nd (FeO.8 Coo. 12 BO.08 ) 5 5
Nd(Fe08CooI2Bo08)55
plus 45wt. % a-Feex-Fe was possibly attributed to the decrease in the crystallization
temperature of the amorphous phase due to the introduction of Co (Zhang
et al. , 2000a).
Lewis and Panchanathan (1999) investigated the relationships between
composition and the nature of metastable phase formation in the low-boron
content melt-spun nanocomposite alloy Nd2[COO 06 (Fe,- xCr x )094
(Fe,-xCrx J232 B
)0 94J232 <
48 (0 ::(
Bl148
<
X ::( O. 09). In general, a high boron to rare-earth ratio, from 1 to 4,
promotes the initial formation of Fe3 B and complex boron-rich intermediate
phases such as cubic Nd2 Fe23 B3 , the hexagonal NdFe'2 NdFe12 B6 and the cubic
Y3 Fe62 Bwtype
B 14 -type phases in the quenched material, while a low compositional
ratio (B / R < 0.5) favors the formation of ex-Fe a-Fe and a 2-17-type
2-17 -type phase.
Chiriac et al. (1999) investigated the magnetic behavior of some
Nd2Fe14 B : a-Fe
ex-Fe nanocomposite alloys with multiple additions of Cu, Nb, Ga
and Co upon some special heat treatment. It has been found that the magnetic-
field annealing enhances the coercivity of the samples with Co addition and a
subsequent stress relief treatment on the same samples improves the exchange
coupling between the grains and hence the remanence and coercivity. A rapid
cooling after annealing improves the exchange interaction whereas flash
annealing leads to the formation of poorly coupled crystallites.
Kobayashi et al. (2000) showed that the high remanence and coercivity
of the Nd-Fe-B : ex-Fe a-Fe nanocomposite alloys containing Nb and Co might
originate from the magnetic coupling between grains via an intergranular
238 W. Liu et al.

ferromagnetic phase. Cui et al. (2000a) investigated the effects of the


addition of Mo alone and the addition of Mo plus Co on the structure and
magnetic properties of mechanically alloyed Nds4 FeS71 Fes7 I B45 . The addition of
Mo alone, especially for the 1.6 at. % Mo case, results in the improvement of
the intrinsic coercivity and the maximum magnetic energy product (BH) (BH ) max'
max .

The addition of Mo alone results in refinement of ex-Fe grains and causes


coarsening of the hard grains. The combined addition of both Co and Mo can
refine the soft grains and inhibit the trend of coarsening of the hard grains due
to the addition of Mo alone. The best magnetic properties J.10M J..IOM He = O. 58 T,
M, = 1.04 T, and (BH)max
J..I oM,=1.04T,
J.1o (BH)max=113.8kJ/m 3
3
= 113.8 kJ/m ,, M,/M
M,/ M s =0. 68, are achieved for
=0.68,
Nds.4Feso.4 C05 . 6MOl IB
Nd84Feso4C05aMol I B45 .
Ji et al. (2000) showed that a 1 at. % indium addition not only increases
the coercivity and remanence of the Nd lo lO Fes3 B 61n nanocomposite, but also
Baln
significantly improves the squareness of the hysteresis loop. The coercivity
and the reduced remanence were increased from O. 50 and O. 71 T to O. 58 and
0.83 T, respectively, and the maximum energy product increases remarkably
from 96 to 143 kJ/m 3 , compared to Nd 10 lO Fes4 B6 a magnets. An excess addition
deteriorates the magnetic properties. The origin for those enhancements is
mainly attributed to the magnetically softened grain boundaries and enhanced
crystallographic coherency by indium addition. Cui et al. (2000b, 2000c)
studied the effects of a partial substitution of Mo, Ga, Si and AI, for Fe on the
structure and magnetic properties of Nd84 Fes71 B45 alloy prepared by
s 4Fes71
mechanical alloying. The addition of 1. 1 at. % Mo, Si or Ga results in the
improvement of the intrinsic coercivity of the nanocomposite magnets. In
contrast, the substitution of 1. 1 at. % AI is unfavorable for improving the
magnetic properties. The optimum average grain size d op of the ex-Fe is
different for various element substitutions; the main effects of these
substitutions on the magnetic properties of the nanocomposite magnets depend
on the grain sizes of the soft- and hard-magnetic phases.
Jurczyk (Jurczyk, 2000; Jurczyk and Jakubowicz, 2000; Jakubowicz
et a\.al. 2000; Jakubowicz, 2001) studied nanocomposite two-phase
Nd2(Fe, Co, Zr) 14 B : ex-Fe
(Fe,Co,Zr)14B magnets prepared by hot pressing of
NdI2.6Fe698-xColl.6ZrxB6 : ex-Fe powder containing 0, 10, 20, 37.5 or 50
Ndl26Fea9S-xCol16ZrxBa
vol. % ex-Fe. It was shown that partial replacement of Fe by zirconium in the
above samples substantially reduces the temperature coefficients of
remanence ex(M,) ex(M r ) and coercivity f3(MH c ) in comparison to sintered Nd-Fe-B.
Generally, if the content of the soft magnetic ex-Fe phase in Ndz N~ (Fe, CoZr) 14 B :
ex-Fe composites increases, the thermal stability of the coercivity increases
too. For the NdI2.6Fe693Coll.6ZrosB6
Ndl2aFea93Co11 aZrQ5Ba : ex-Fe magnet, containing 37.5 vol. %
ex-Fe, the temperature coefficients (from 293 to 413 K) of remanence ex and
coercivity f3 are - 0.07 and - O. 35 % K -I , respectively.
A comparative study of mechanically alloyed and mechanically milled
Nd 10 lO Fes4 B6
a has been made by Miao et ai.,al., (1996). The results indicate that
the mean grain size of mechanically milled powders was considerably smaller
Nanostructured Exchange-Coupled Magnets 239

than that of mechanically alloyed powders. Moreover, the grain size


distribution of the mechanically milled powders was more uniform. As a
consequence, both the remanence and coercivity of mechanically milled
powders were superior to those of mechanically alloyed powders. A similar
conclusion has also been reached by Cui et al. (2001) in studying the structure
and magnetic properties of Nd84 84 Fe86 MOll B45 nanostructured magnets
prepared by mechanical alloying, compared with those by mechanical milling.
Nanocomposite magnets appear to be more corrosion resistant than
sintered Nd-Fe-B magnets. For the Nd l26 Fe668 C0 CO 116
l16 All Crz
Cr2 B6 : 37. 5 vol. %
a-Fe
ex-Fe magnet, the corrosion rate was 0.037 mm/year. Effective protection
against corrosion for nanocomposite Nd-Fe-Co-AI-Cr-B : a-Fe ex-Fe magnets was
realized by surface coating with Zn. You et al. (2000) investigated the effect
of Co addition combined with W on the microstructure and magnetic properties
of nanocomposite magnets of Nd855 Fe8449 WO. 60 BO. 636 636 and Co prepared by
mechanical alloying. W addition reduces the grain size of the soft phase, while
Co addition reduces the crystallization temperature of the hard phase.
Neu and Schultz (2001) studied exchange-coupled Nd2 z Fel4 B : a-Fe
ex-Fe
powders, prepared by mechanical milling and subsequent heat treatment. The
saturation polarization and coupling strength increase with Co substitution
(deduced from an enhancement of M,/M Mr/M ss )'
)' The Tc c measurement for a series
of Co-containing samples reveals an identical Fe : Co ratio in both phases,
which is controlled by the ratio chosen for the initial composition. The
decrease in the anisotropy constant of the 2 : 14 : 1 phase imposes a practical
upper limit for the Co substitution, exhibiting a decreased coercivity for the
samples with Co content above 10 at. %. The improvement based on Co-
substitution is Iinked
linked to a homogeneous and fine-grained microstructure,
resulting from small additions of Si or Zr and the use of a stoichiometric master
alloy. The mechanically milled and optimally annealed Nd-Fe-B powders show
high energy products of (BH)max~ 150 kJ/m 3 3
for a phase ratio of 70 vol. %
hard and 30 vol. % soft magnetic phase. The magnetic properties, processing
of Nd2zFel4
Fe14 B : ex-Fe
a-Fe type nanocomposite magnets with different compositions
are listed in Table 6.9.
According to the results mentioned above, it can be seen that the
maximum energy product has not been significantly improved by adding some
3d elements into the Nd2 z Fel4 B phase in exchange coupled nanocomposite
magnets, although coercivity or Curie temperature is somewhat enhanced.

6.3.3.6 Combination Substitution for Nd and Fe in


Nd 2 Fe J4
l4 B:
B : a-Fe Nanocomposites

Betancourt and Davies (1999) observed an enhancement of remanence and


magnetic energy product for didymium-iron-boron, Nd2 z- xx Pr x Fe14 B, with Nd :
Pr ratios of 1 : 3, 1 : 1, and 3 : 1 and for (Nd, Pr)yFe94-yB6, with y=8, 10,
12 and 14. This enhancemnt arises from intergranular exchange coupl coupling
ing
between the Rz2Fel4 B crystall
crystallites
ites which have a mean diameter of about 40 nm
240 W. Liu et al.

Table 6.9 110 M" intrinsic coercivity, 110M


Remanence, fJoM" He'
fJoMH MJ Ms and
c ' remanence ratio, M,/M
maximum energy product (BH)CBH )ma,
max of Nd2 Fe,.
Fel4 8
B : a-Fe
ex-Fe type nanocompsite magnets with
different compositions, with different synthesis and with different 3d element additions.
M,
fJoM,
110 110M
fJoMHHec CBH)ma,
(BH)max
Compositions Synthesis M,/ M s Reference
(T) (T) CkJ/m 3 )
(kJ/m
Mechanical
07
.07 0.40 0.66 92 [lJ
alloying
Mechanical
NdsFes7
FeS7 Zr184
Zr, B. .04 043
0.43 0.65
065 96 [lJ
[1J
alloying
Mechanical
Nd9.5 Fess.s Si, 8 45 099 0.47 109 [2J
alloying
Mechanical
1.0 0.53 121 [2J
alloying
Nd,o
Nd Fes4 B
lO Fes, 86 Rapid quenching 0.50 0.71 96 [3J
Nd,o Fes3 8 61n
FeS3 B Rapid quenching 0.58
058 083 143 [3J
Nd7Fes2.sC06.sZro.s83.s
Nd7Fes2.5Co6.5Zro.s B3.S Mechanical milling 1.21
121 042 0.70 129 [4J
NdasFe7s
Nd Fe7S C09
COg Zr, B.
84 Mechanical milling 112
1.12 043
0.43 0.68 117 [4J
Nds Fe7S C09
COg Si, B,
84 Mechanical milling 1.19 0.50 0.72 149 [4J
Ndg Fen C0
Coss Si 1 B
85 Si, 8 45
u Mechanical milling 1.11
111 0.60
060 0.70 147 [4J
Nd lO Fe
Nd,o 74 Co
Fe74 COlO Sh B
lO Si, 8s Mechanical milling 103
1.03 0.61 0.69 128 [4J
Nd(Feo sCoo 12
NdCFeo SCoo os )s s +
8 0 os)s
'2 Bo
.. .. . . Mechanical milling 042
0.42 071
0.71 142 [5J
45%
NdsFe7sCosSi,84
NdsFe7sCOgSi,B, Mechanical milling 1.19 0.50 0.72 149 [4J
Nds.4Fes6.sMo06845
Nd84 Fes6.5Moo.6 Bu Mechanical milling 0.99 0.40 0.67 94 [6J
Nd84
s.4Fess.s M0
MO l6 8u
16 B 45 Mechanical milling 0.95 0.50 0.64 93 [6J
Nd84Feso.4Cos.6Mo11845 Mechanicalmilling
Nds.,FesQ.4COS6Mo"Bu Mechanical milling 1.04 0.58 0.68 114 [6J
Nd84 FeS6Ga 11 8 45
Nds,FeS6GaIlB,s Mechanicalmilling
Mechanical milling 1.07 0.40 0.68 93 [7]
Nd84
s , FeS6 Al
AlII 8u
l1 B 45 Mechanical milling 0.98 0.30 0.64 73 [7]
Nels FeS7 Nb,
Nd;; 86
Nb l B Rapid quenching 1.04 0.38 0.67 78 [8J
Nd 4Feso
Nd, Fe90 Nb28
B,4 Rapid quenching 1.45 0.27 0.79 115 [8J
Nd4FesoNb284
Nd,Fe90Nb2B, Rapid quenching 1.350.270.88 103 [9J
Nd6Fe
Fess 86
ss B Rapid quenching 1.31 0.34 0.71 76 [10J
[1 OJ
Nd6FessTi,8
FessTb B6 Rapid quenching 1.010.41074
1.010410.74 83 [10J
Nd6FeS3 Ti
Nd6FeS3 TissB
86 Rapid quenching 9.40 0.45 0.71 76 [1 OJ
[10J
Nd6Fess V33B
86 Rapid quenching 1.30 0.32 0 80
0.80 104 [10J
[1 OJ
Nd
Nd;;6FeS6 CuNb8
CuNbB 6 Rapid quenching 1. 32 0.31 0.77 103 [11 J
[1] Neu et al.,
[lJ aI., 1996; [2J Wecker et aI.,
al., 1995; [3J Ji et aI., 2000; [4J Neu and Schultz, 2001;
[5J Zhang, 2000a; [6J Cui et al.,
aI., 2000a; [7J Cui et al..
al. , 2000b; [8J Hadjipanayis et al..
aI., 1995;
[9J Konno et al..
aI., 1998; [10J Fang and Chin, 1996; [11J
[llJ Fang et al..
aI., 1999.

and also between the R2 Fe,. ex-Fe grains. For y = 12, the lJo
Fe14 B and a-Fe J..lo M r
enhancement can be attributed to the exchange coupling between the R2 Fe14
Fe,. B
Nanostructured Exchange-Coupled Magnets 241

crystallites with fine grain size. The progressive increases of J..lo Jio M, between y
= 8 and 12 is due to additional coupling between the R2Fe14B
=8 R2Fe14 B and a-Feex-Fe grains.
On the other hand, the decrease of JiJ..l 0M, 0 M, between y = 12 and 14 is due to the

presence of a paramagnetic rare-earth-rich phase around 2 : 14 : 1 grain


boundaries, which both decouples and dilutes the magnetic grains. An
excellent combination of (BH)max(185 kJ/m 3) and intrinsic coercivity (1.0 T)
was achieved for (Pro(Pr07sNd02s)2FeI4B.
75 Nd025 ) 2Fe14 B.
Zhang et al. (2001 c) investigated microstructure and magnetic properties
of melt-spun nanocomposite magnets with nominal compositions of (Nd 1l - xx
Pr x)9 Fes6
FeB6 B5 = 0 - 1). Substitution of Pr for Nd could significantly improve
s (x =
the hard magnetic properties of the nanocomposite magnets; the intrinsic
coercivity and the maximum energy proooct product increase from 0.52 T and 124 kJ/ m3 for
3
x = 0 to 0.62 T and 152 kJ/ m for x = O. 6, respectively. Further substituting Nd
>
by Pr (x > O. 6) strongly weakens the exchange coupling between
magnetically hard and soft phases. According to the exchange-coupling model
of Kneller and Hawig (1991), to obtain a sufficiently strong exchange
coupling, the optimum grain size of the soft magnetic phase should be
approximately equal to the Bloch-wall width of the hard phase. From the
experimental results, it is concluded that a large amount of the substitution of
Pr for Nd may greatly increase the first-order magnetocrystalline anisotropy
constant of the hard magnetic phase, and decrease the optimum grain size of
soft magnetic phase, resulting in a weaker exchange interaction between the
hard and soft phases in comparison to Nd9 Fes6 FeB6 B5s . It is deduced that the
increase of energy product with increasing Pr content (x < o. O. 6) results from
the improvement of the rectangularity of the demagnetization curve, and the
>
decrease of energy product with more content of Pr (x > 0.6) arises from the
decrease of the remanence. The melt-spun amorphous (Nd, Pr)yFe93-X-YCOX
(Nd,Pr)yFe93-x-yCoX
Nb2B5 s(x = 0 - 20, y = 5 - 7 at. %) ribbons in the as-quenched state changes
to a nanocomposite structure consisting of a-(Fe, ex-(Fe, Co), (Nd,Pr)2(Fe,Co)14B
and residual amorphous phases after annealing at temperatures of 973 -
1023 K. The magnetic properties are improved after anneal ing at a higher
heating rate in the temperature range corresponding to the primary
crystallization temperature of a-Fe ex-Fe phase (Kojima et al. , 2000).
The magnetic properties of (Nd,DY)2 (Fe,Nb)14B : a-Fe ex-Fe nanocrystalline
composite, containing 0 to 30 wt. % a-Fe, ex-Fe, prepared by melt spinning and
anneal ing have been studied by Wang et al. (1997). The best
subsequent annealing
values of remanence, coercivity, and maximum energy product (BH) max are
1.02 T, 0.88 T and 134 kJ/m 3 , respectively, for NdB s.,16
16 DYI FeBs26
Fe8526 Nb 1B
BUB
U8 .
microstr~cture
The microstructure of the nanocomposite consists of hard magnetic (Nd, DY)2
(Fe, Nb) 1414 B and soft magnetic ex-Fea-Fe with an average size of about 30 nm.
Chang et al. (1998) obtained a coercivity of more than 1. 1 T and an
energy product of more than 127.4 kJ/m 3 for melt spun (Nd09sLaooshs+xFeba'
(Ndo.95Lao05)75+xFebal
Cr2B
Cr2 10 powders, where x ranged from 3 to 3. 5. It was found that a slight
B 10 sl ight
substitution of Cr for Fe suppresses the formation of the R2 Fe23 B3 and Fe3 B
242 W. Liu et al.

phases during crystallization and results in a R2 z Fe


Fe1414 B : ex-Fe mixture, and
increasing the total rare earth content leads to enhancement of the magnetic
properties of the alloy powders. The magnetic properties including remanence
of 0.92 - 0.97 T, intrinsic coercivity of 1. 11 - 1.32 1. 32 T, and maximum magnetic
energy product of 131. 4 - 143. 3 kJ/m 3 have been obtained for (Ndo .9s
095
Laooshs+xFeBo.s-xCrzBlO(x = 3 - 3 . 5).
Lao 05) 75 + x FeS05- x Cr 2 B ,o (x =3-3. 5) .
Chen et al. (1999b) and Lu et al. (1999a, 1999b) investigated the phase
transformation, microstructure and magnetic properties of (Nd oo95 Lao. 05)95
.9s Lao OS)95 Fe bal

CO
C05 Nb
s 2 B
z lO .
lO 5
. s nanocomposite. The Henkel plot (Henkel, 1964) was employed to
study the strength of the exchange coupling between the hard and soft magnetic
phases in the as-spun and thermally treated samples. The crushed sintered Nd-
Fe-B showed mostly negative 8M, oM, suggesting a magnetostatic-dominated
inter-particle interaction. The 8M oM of the as-melt-spun materials and samples
annealed at 650 'c exhibited similar behavior. For samples heat-treated at
above 650 'c, a positive oM 8M was observed indicating the existence of
exchange coupling between phases. The height of the 8M oM peak reached a
maximum when materials were treated at 700 'c , suggesting that the strongest
exchange coupling existed among samples studied. Annealing at 750 and 850 'c ,
the peaks of oM 8M decreased significantly, implying a weakened exchange
interaction. The Henkel plot successfully interpreted the effect of the exchange
coupling on 110 IJo Mr.M He and (BH ))max max obtained for samples annealed below

750 'C. Excessive grain growth producing grains from 38 to 68 nm in diameter


was observed in the ribbon annealed at 850 'C. Soft magnetic phases such as
Fe3 B and ex-Fe precipitated at the grain boundaries. These intergranular
phases are exchange coupled with the hard phase causing a decrease of M He.
They also reported types of magnetic domain patterns caused by either dipolar
coupling or exchange coupling inside a heat-treated boron-rich Nd-Fe-B melt-
spun ribbon. Single particle domains were commonly observed in Fresnel
images of all the samples, suggesting little exchange coupling among the
grains in these regions. Snake-shaped long interactive domains consisting of
chains were found in many Foucault images at the regions where exchange
coupling is weak. These domains are caused by dipolar coupling along the
chains. Wider domains consisting of many grains were also observed by
Fresnel images, although they were much less common than the single particle
domains. The grains inside a wider domain are believed to be exchange
coupled with each other. The magnetic properties of such boron-rich Nd-Fe-B
composites are influenced more by the dipolar coupling than by exchange
coupling
coupl ing between the crystalline
crystall ine grains.
6.3.3.7 Pr2 Fe14
I4 B: a-Fe Nanocomposite Magnets

The intrinsic magnetic properties of Prz Fe14 B are similar to those of Nd2Fe,4B,
Pr2Fe14B NdzFe14 B,
but Prz Fe14 B can be used at low temperature, because there is no spin
Pr2 Fel4
reorientation in Prz Fe14 B. Therefore, Prz
Pr2 Fel4 Fe14 B is also a good candidate for
Pr2 Fel4
hard phase in nanocomposite magnets.
Nanostructured Exchange-Coupled Magnets 243

Villas-Boas et al. (1997) studied flash annealing and magnetic


interactions in Pr4 Fe78 B 18 prepared by melt spinning. M6ssbauer spectra
showed that flash annealing resulted in significant amounts of Pr2 Fe23 B3 and
reduced quantities of ex-Fe. a-Fe. The optimum magnetic properties of 110M =
POM He =
M, / M
O. 35 T, M,/ Mss = kJI m3 were achieved in the material
= 0O.. 77, and (BH) max == 88 kJ/
annealed at 600600CC for20 s.
Murakami et al. (2000) produced ribbons by melt spinning and
subsequent flash or conventional annealing to develop nanocomposite magnets
with enhanced remanence. The aim of their work was to examine the magnetic
properties and microstructure of the following three Pr-Fe-B samples with
different annealing
anneal ing treatments: (1) the sample flash annealed with an electric
current density j = 1. 90 x 10 7 A/m AI m2 that showed the highest coercivity
exhib iting good coupling;
exhibiting coupl ing; (2)( 2) the sample flash annealed with j = 1. 68 x
10 77A/m
AI m2 that showed partial coupling;
coupl ing; and (3) the sample conventionally
annealed at 660C that showed the highest remanence ratio M, M,/I Ms. Results
show that the phases present in the samples flash annealed with j = = 1.90 X 107A/m
A/m22
and conventionally annealed are Pr2 Fe14 B, Fe3 B, a-Fe, ex-Fe, while in the sample
7 2
2
flash annealed with j = 1. 68 XX 10 A/m the Pr2 Fe23 B phase is found. A
Rietveld analysis of X-ray diffraction and neutron diffraction data showed that
the main phases in the three samples were the same: Pr2 Fe14 Band
Pr2Fe14B Fe3 B. No
and Fe3B.
amorphous phase was detected.
Goll et al. (1998) carried out magnetic investigations of rapidly quenched
exchange coupled single-phase and composite magnets containing Pr2 Fe14 B
ex-Fe ranging from 0 to 56. 1 vol. % in the temperature region 20 K~ T ~
and a-Fe
560 K. In Fig. 6.38, the room temperature hysteresis loops of different
compositions Pr2 Fe14 B-based magnets consisting of optimally quenched ribbon
flakes are compared with one another. The exchanged coupled samples show
a significant remanence enhancement in comparison with decoupled Pr15 Pr,sFe78B7
Fe78 B7
magnet with over-stoichiometric Pr (PoMr~ (l1oM,~ 0.511 =
O. 5P 0oM s =0.78
O. 78 T). Thus, the
remanence of the stoichiometric sample Pr'2 Pr12 Fe82 B6 (110
(Po M,
Mr == 0.95 T) is
enhanced by 21. 1 1%.
%. In the exchange coupled composite magnets a further
enhancement of the remanence up to a maximum value of Po 110 M, == 1.42 T in
Pr6FegoB4, containing 46.9 vol. % a-Fe,
Pr6Fe90B4' ex-Fe, was observed. As the hysteresis
loop of the stoichiometric sample is nearly rectangular, the squareness S is
very high (S = O. 71). S is defined here by the maximum of the product MH
along the demagnetization curve divided by M, He. With increasing a-Fe ex-Fe
content this rectangularity deteriorates so that S becomes smaller. This is the
reason why the remanence does not increase as much as expected with
ex-Fe content. However, the squareness of the hysteresis loop is
increasing a-Fe
improved by the exchange-coupling phenomenon at least for samples
containing not more than 46.9 vol. % a-Fe. ex-Fe. In Fig. 6.39 the room temperature
magnetic properties of all investigated exchange coupled Pr2 Fe14 B-based
magnets are summarized as a function of the a-Fe ex-Fe concentration. The best
room temperature magnetic properties 110M POM He == 0.59 T, Po 110 M, == 1. 17 T and
244 W. Liu et al.

(BH)max =
( BH) max = 180.
180.77 kJ/
kJ/m 3
m are obtained in Pr 30. 4 % ex-Fe. As at low
s FeS? 8 5 with 30.4%
PrsFes785
ex-Fe concentrations 30 vol. %) the coercive field is higher than one half of
the remanence, (BH) max sensitively depends on the remanence according to
theoretical upper limit of isotropic samples (BH)max~J.1oM~/4.
(BH)max~j.JoM~/4. If at high ex-Fe
concentrations (> 30 vol. %) the coercive field is smaller than one half of the
remanence, (BH(BH)) max is Iimited
limited by irreversible demagnetization processes
leading to a drastic decrease of (BH)
(BH)max'
max'

J..-
f-"" ...--
-:.-
.

1--7
"-7 .//" tl-
Decoupl~ _ ~I-
Decoupled
-J
-I- r---~
r..--/ //'
,

I
.--- ~~

I
I/ ---- I

JI
I
-_/ --'
I/

r..--/
1--/
~
.;;;/ LJ-
- -1- l-.t ~
'-J.
IH- I~ -
l- Stoichiometric
.L--11
i==-
-I F-
=-----
..-/ -1- 1:1. 30.4% a-Fe
:::....- --- 46.9% a-Fe
-2
----- o -- 2
llo H(T)
floH(T)

Figure 6. 38 Composition of room temperature hysteresis loops between different PrFeB-


based magnets: Pr,5
Pr'5 Fe78 B7(decoupled), Pr'2
Pr12 Fe82
Fe" B6 (stoichiometric), Pr8 Fe87 B5 (30.4 %
Pr6 Fe90 B, (46. 9% a-Fe) (Goll et ai.,
a-Fe) and Pr6FegoB,C46.9% aI., 1998).

2.0 lloHc>0.5 if
floHc>0.5 Jf floHc<0.5 Jf
lloHc<O.5 if 200
__ I
_I
1.5 (BH)max
(BH)max I1 150 ~

b:
...,
'-)
I1
~:..,
g
E 100 6
g
:r:::t:
uu

~
::f
of
0.5 50 ~~
0.0
0 10 20 30 40 50
a-Fe concentration (vol. %)

Figure 6. 39 Magnetic properties at room temperature as a function of the a-Fe


al. , 1998).
concentration of composite PrFeB-based magnets (Goll et ai.,

Mendoza-Suarez and Davies (1998) showed that the mean Pr2 Fe'4 8
crystallite size within the nanocrystalline range had a marked effect on the
magnetic properties for melt-spun Pr x Fe94- x 8 6 alloys (with 6 ~ x ~
comb inations of M He
20 at. %). Excellent combinations He and (BH) max were observed for Pr in

the range 9 - 12 at. %, and the system appears to offer clear advantage over
corresponding nanophase NdFe8 alloys in having significantly higher M He for a
given enhancement of (BH) max' Wang et al. ( 1999, 2000a, 2000b) found that
Nanostructured Exchange-Coupled Magnets 245

during annealing treatment of Pr7 Fess Feaa 8 B5 ribbons prepared by melt spinning at
lower wheel speed (22 m/ ml s or 26 m/s), ml s), the amorphous phase in the as-
quenched ribbons transforms directly into Pr2 Fel4 Fe14 B
8 and ex-Fe phases.
However, both the metastable 1 : 7 and Pr2Fe14 Pr2 Fel4 B 8 phases form for the case of
velocity v = 30 m/s, mis, and only the metastable 1 : 7 phase occurs in the ribbon
melt spun at v v=34= 34 m/s. In PrsFes686
PraFea6B6 ribbons, the amorphous phase in the as-
quenched ribbons transforms directly into Pr2 Fel4 Fe14 B 8 and ex-Fe. By 8y contrast,
after initial crystall ization of ex-Fe but prior to the formation of the final mixture
crystallization
of Pr2 FeFel4 l 48
B and ex-Fe phases, both the metastable Pr2 Fe23 B 8 3 and Pr2 Fel4 8 B
phases form for the case of a wheel velocity v = 26 m/s, mis, and only metastable
Pr2Fe2383
Pr2Fe23B3 phase occurs in the ribbon melt spun at v=30 v = 30 m/s.
Chen et al. (1999c) developed a uniform Pr2 Fel4 Fe14 B8 : ex-Fe nanocomposite
structure with fine ex-Fe grains in the melt-spun PrsFes6 Pra Fea6 B8 6 ribbons. Lower
speed leads to large 2 : 14 : 1 and ex-Fe grins, while higher speed leads to the
appearance of an amorphous phase that will result in a Pr2 Fel4 Fe14 B
8 : ex-Fe
structure with large ex-Fe grains after subsequent crystallization annealing.
Mendoza-Suarez et al. (2000) ( 2000) experimentally studied correlations of
J.1 o M, and (BH)max for a series of Pr
MH e , 110M, x Fe94-x 8 6 , with x ranging from 6
PrxFe94-xB6'
at. % to 20 at. %. The results showed that the variation of the Pr content and
the casting parameters yielded to 110MH J.1 0MHee ranging from values near 0 T to about
2.8
2. 8 T. 110 J.1o M, increased at the expense of 110M J.10M He in agreement with results
reported by Goll et al. (1998). (BH) max increases with 110 J.1o M, up to about 30 %
ex-Fe content, beyond this point (BH) max fell due to its strong dependence on
110M
J.10M He' and 110 J.1o M, increased from about 0.6 to 1. 1 T as MHe decreased from
about 2. 8 T to about o. O. 25 T. On the other hand, (BH) max increased for values
J.1oMH e up to 0.57 T and then decreased for higher(l1oMH
of 110MH higher(J.1oMH e ) values.
Liu et al. (1999c) systematically studied the structure and magnetic
properties of (Nd, Pr) -Fe-8-Ti -Fe-B-Ti alloys prepared by mechanical alloying. The
best magnetic properties, 110M, =
J.1o M, = 0.8 T, 110M =
J.10M He = 1.9 T, M,/ M,/M M s..' and
(BH)max = 111.2 kJ/m 3 3
,, were obtained in PrI6Fe76Ba
Pr16Fe768s For PrlOFeS28s,
Pr10Fea2Ba, it was
found that the remanence ratio M,/ M,I M s and the maximum energy product
decrease with increasing grain size. The action and occupation state of Ti in
Prlo Fe76 8 Ba s- z Ti 6+ z alloys for 0 ,s;;; ~ z ~ ,s;;; 6 are very different from those in the
Nd-containing counterparts. Harland and Davies (2000) showed that
substitution of 1 at. % Zr for Fe in Prl0 Prlo Fes
Fea44 B8 6 resulted in an improved loop
shape and 110M J.1o MHe is in the range 0.66 O. 66 - 0.69 O. 69 T, and (BH) (BH ) max increased to
130-140 kJ/m 3 3
as compared with 80-100 kJ/m 3 3
for Zr-free alloys.
Chen et al. (2001 a) conducted a comprehensive study on the effect of
various substitutions on the magnetic and structural properties in melt-spun
Prs
PraFea4M2B6
Fes4 M28 6 (M = (M=Cr,Cr, Nb, Ti, and Zr). Nanostructure studies revealed that M
substitutions yield a finer and more uniform 2 : 14 : 1 : ex-Fe nanostructure.
The largest improvement of the magnetic properties, as compared to Prs Pra Fee6
Fea6 B8 6 is
obtained in Nb substituted Prs Pra Fes4
Fea4 Nb 2 8
B
2 6 magnets where a coercivity of O. 65 T
and a maximum energy product of 143 kJ/m kJI m3 have been obtained. Zhang et al.
246 W. Liu et al.

(2001 d) investigated the phase evolution and magnetic properties of melt-spun


Pr9FeS6-xCoxB5(X=0, 6,10,12,14,16)
PrgFes6-xCoxBs(x=0, 6,10,12,14, 16) nanocomposites. It was found that
substitution of Co for Fe raises the magnetization under an applied field /-10 /-10 H =
1.50 T for x=O - 1.80Tforx=12.
6.5T, from 1.50Tforx=0 1.80 T for x= 12. While/-loM"
While /-10M" /-10M
/-10M He and
( BH) max changes from 0.96
(BH)max O. 96 T, 0.57
O. 57 T and 94 kJ/ kJ/mm3 for x =0= 0 - 1. 15 T, 0.63
1.15 O. 63
137. 7 kJ/ m3 for x = 12, respectively.
T and 137.7
Chen et al. (2001 b) investigated the enhancement of magnetic properties
of nanocomposite Prs-xDYxFes6B6(x=0,
ofnanocomposite O. 5, 1, 1.5
Prs- x Dy x FeS6 B6 (x = 0, 0.5,1, 1. 5 and 2) magnets. An
optimum coercivity of O. 66 T is obtained in the Pr7 Dy, FeS6 B6 magnet as
DYI Fes6
compared to 0.43 T in the Prs Fes6 FeS6 B6 magnet. As a result, the energy product
is increased from 72 kJ/m 3 3
in the Prs Fes6FeS6 B6 magnet to 135 kJ/m 3 in the
FeS6 B6 magnet. Microstructure studies reveal a finer and more uniform
Pr7 DYI Fes6
2 : 14 : 1 : a-Fe
ex-Fe nanoscale microstructure in the Dy substituted samples, due
to the low wheel speed where the microstructure is crystallized directly from
the melt. The enhancement of magnetic properties by Dy substitution is mainly
due to the nanostructure refinement that results in an enhanced exchange
coupling between the Pr2 Pr2Fel4B a-Fe, whereas the anisotropy increase by
Fe 14 B and ex-Fe,
the Dy substitution provides only minor contribution to the enhanced coercivity.
Withanawasam et al. (1996b) studied a new nanocomposite magnet consisting
of a mixture containing a magnetically hard Pr2 C0 CO 14
l4 B phase and a soft Co
phase. The final metastable TbCuy-type
TbCu7 -type phase together with the above phases
is formed after annealing the melt-spun ribbons through an intermediate
transformation. An optimum coercivity of O. 43 T was observed in PrSCO , Bs .
Prs COS3 Nb 1
O'Sullivan et al. (1999) prepared a Pru5Tb1Fe71C025Cr3B,s
PrusTblFe71Co25Cr3B1S composite
by mechanical milling and subsequent annealing. The material contains a
(Pr, Tb)2(Fe, CO)14B hard phase intimately mixed with (Fe, Cr)2B and a-Fe ex-Fe
soft phases. Wang et al. ((2000c) 2000c) investigated the phase composition,
microstructures and magnetic properties of melt-spun Prs.s Prs 5 (Feu
(Feo SCOCOOO.22)) S65- x
CuxB O. 5, 1.0,
CU xBs5 (x = 0, 0.5, 1. 0, 2.0,
2. 0, 3.0)
3. 0) nanocomposites. It was found that Cu
addition suppresses the formation of the Pr2 (Co, Fe) 17 phase and results in a
two-phase mixture of a-(Fe,ex-(Fe, Co) and Pr2(Fe, CO)14B. Transmission electron
microscopeCTEM) and energy dispersive X-ray(EDX) X-ray (EDX) analysis show that Cu
strongly segregates at intergranular regions. The magnetic properties of the
ribbons initially increase with increasing Cu content, and reach maximum
values at O. 5 at. % Cu, and then decrease for higher Cu concentration. The
variation of the magnetic properties with increasing Cu content can be
attributed to the variation of the exchange coupling between the hard and soft
phases and the decrease in saturation polarization of the ribbons. A
/-10M He = 0.74 T, /-10
combination of /-10M M r = 1.09 T and (BH
/-10 M, ) max = 160 kJ/m 3 has
(BH)max
been obtained in a PrS5 Prs.s (Feu CO COOO.22)S6 CUO.5 B5
)S6 Cuo.s s ribbon. Moreover, the minor
addition of Cu (x = O. 5) is effective in reducing the irreversible loss of
induction of the studied Pr2(Fe, CO)14B/ a-(Fe, ex-(Fe, Co) nanocomposite magnets.
Wang et al. (2000d) reported the high-performance Pr2 Fel4 Fe14 B/ex-Fe
B/a-Fe
nanocomposite magnets prepared by hot pressing melt spun Prs Dy, DYI Fe745
Fe74S COlO
Nanostructured Exchange-Coupled Magnets 247

Nb05 B6 flakes under a conventional pressure of 125 Mpa and high pressures
ranging from 1 to 7 Gpa. It is evident that the grain sizes of both 2 : 14 : 1
phase and ex-(Fe, Co) decrease with the increase of compaction pressure.
When increasing pressure from 125 Mpa to 5 Gpa, the grain size of 2 : 14 : 1
phase decreases from 22.5 to 8.7 nm, while the grain size of ex-(Fe, Co)
phase also drops from 33.3 to 10. 1 nm. A remanence of 1. 11 T, coercivity of
1. 0 T, and maximum energy product of 188 kJfm 3 have been achieved in the
hot pressed magnet under a pressure of 5 Gpa. Processing and magnetic
properties of Prz
Pr2 Fe!.
Fe l4 Bfex-Fe type nanocomposite magnets with different
compositions are concluded in Table 6. 10.

/Jo M" intrinsic coercivity, lJoMH


Table 6.10 Remanence, lJoM" M,/ Ms and
e , remanence ratio M./M
/JOM He'
maximum energy product (BH ))ma,max of Prz
Pr, Fel.
Fe" 8/ex-Fe
B/a-Fe type nanocomposite magnets with
different compositions and with different synthesis. All samples are produced by rapid
quenching.
IJOM
/JoMHHee (BH)ma,
(BH)max
Compositions lJo
/Jo M, (T) M,M s
M.M Reference
(T) (kJ/m 3 )
Pr, Fe7s
Fe78 8
B,s
18 1. 15
1.15 0.35 0.77 88 [lJ
PrlS Fe78 B
PrIS Fe7s 8 7 (decoupled) 0.75 2.0 100 [2J
Pr12 Fes,
Pr" Fe82 B
8 6 (stoichiometric) 0.95 1.2 156 [2J
Prs Fes7 Bs (30.
Pr6Fe878s (30.44 % ex-Fe)
a-Fe) 1.17
1. 17 0.59 181 [2J
Pr6 Fego 8,
Pr6FegoB, (46.
(46.11% ex-Fe)
a-Fe) 1.42
1. 42 0.29 132 [2J
Pr8 Fes6 B
Prs 86 1. 09
1.09 0.43 0.65 72 [3J
Prs Fes, Cr,
Cr2 B
86 1.07 0.52 0.72
072 88 [3J
Prs Fes, Nb 86
Nb,2B 1. 08 0.65
065 0.70 143 [3J
Prs Fes, Ti,
Tiz B
86 1.19
1. 19 044 0.72 102 [3J
Prs Fes,
Fes4 Zr2 86
Zr, B 0.93 0.54 0.62 108 [3J
Pr7 Dy,
DYl Fes6
Fe86 8
B6 1.14 0.66 0.71 135 [4J
Prg Fes6 B
8s 0.96 0.57 0.64 94 [5J
Prg Fe7' C012 8s
Co" Bs 1.15 0.63 0.64 138 [5J
PrB.S Fe6g.,
Prs.s CO,7 .33 8
Fe69.2 Co17. Bs 1.03 0.58 0.67 129 [6J
Prs.s Fe68 sC017.2Cuo.
PrBS Fe6BB COil' Cuo.ss8s
Bs 1.09 0.74 0.71 160 [6J
Pr8sFe6s,Co171
PrB5 Fe6B, COil I CUl 8s
CUI B 1.03
1. 03 0.69
069 0.68 138 [6J
Prs DYl Fe7,.sCo
PrB Dy, lO Nbo.s8
Fe7'.s COlO B6 1.11
1. 11 1.02
1. 02 188 [7]
[1] Villas-Boas et at.,
al. , 1997; [2J Gall
Goll et at.,
al. 1998; [3 J Chen et al.
at. , 2001 a; [4 J Chen et at.
al .,
2001b; [5J Zhang et at. at. , 2000c; [7]
aI.,, 2001d; [6J Wang et aI., [7J Wang et aI.,
at. , 2000d.

6.3.3.8 Nd-Fe-B Permanent Magnet Films


Generally, a preferred microstructure of nanocomposites may be more easily
realized through pr~paring nanocomposite film magnets than ribbon or powder
materials, because of the possibility to control structure and composition of the
film at will. Cadieu et al. (1986a, 1986b) first reported highly coercive
248 W. Liu et al.

perpendicular anisotropy Nd-Fe-B films synthesized by selectively thermalized


sputtering onto heated substrates. A remanence of 0.93 O. 93 T and a coercivity of
1.6 T have been achieved at room temperature. Later, Nd-Fe-B films have
been prepared by using thin-fi 1m processes such as magnetron sputtering,
thin-film
molecular beam epitaxy, and pulsed laser deposition (Aylesworth CAylesworth et al. ,
1989; Cadieu, 1987; Kapitanov et al. , 1993; Lemke et al., 1996; Keavney
etal.,
et al. , 1996, 1997).
Araki et al. (1999)
C1999) investigated the texture and magnetic properties of
Nd12Fe88-xBx
Nd 12 FeBB- x B x films prepared by magnetron sputtering. The results showed that
the c-axis orientation of the Nd2Fe14 B phase in the film was higher, and the
perpendicular coercive force became greater with an increment of x.
However, for x > 15 the coercivity decreases due to the formation of a large
amount of an amorphous soft magnetic phase. The transmission electron
microscopy observations revealed that the Nd 12 l2 Fell
Fe71 B 17
17 film consists of
alternating oriented Nd2Fel4 Fe14 B and amorphous phases with a thickness of about
20 and 5 nm, respectively. This material is an example of an exchange-
coupled hard-soft nanostructure with an aligned hard phase.
Parhofer et al. (1998)
C1998) reported that the sputtered Nd 16 '6 .7 Fe7555B
7Fe75 Bu78 thin
films with thickness from 5 nm to 350 nm, deposited directly onto a quartz
substrate, grow with a pronounced c-axis texture perpendicular to the film
plane and have a coercivity of about O. 20 T. If the film thickness is smaller
than about 150 nm, the oxygen in the substrate oxidizes a significant part of
the Nd and the coercivity decreases to values below O. 03 T. The deposition of
an 80 nm Cr buffer layer between the substrate and the Nd-Fe-B film prevents
the oxidation of the films but negatively affects the growth of texture. The Nd-
Fe-B films with a thickness of 20 nm have good, hard magnetic properties
(J.10M Hec = 1.0 T, J.10
ClJoMH M r = 1. 1 T, and a maximum energy product of 190 kJ/m 3
1J0M,=
perpendicular to film plane) .
Jiang et al. investigated Nd 2Feu BO. 8 thin films of the type A (20
Nd2FeuBoB nm) /Nd-
C20 nm)/Nd-
Fe-B(d
Fe-BCd nm) / A(20 nm), where d ranges from 54 to 540 nm and the buffer layer
A is Nb or V. The films were prepared on a Si (1 C1 0 0) substrate by magnetron
sputtering. A 30 s rapid annealing or 20 min conventional annealing forms the
Fe14 B phase. The average crystall ite size ranged from 20 to 35 nm,
hard Nd2Fel4
with the rapidly annealed samples having the smaller crystallite size. The
maximum coercivity of 2.63 T at room temperature was obtained for an Nb/
Nd-Fe-B (180 nm) /Nb film after a rapid annealing at 725 'c
C180 nm)/Nb 'C..

6.3. 3. 9
6.3.3.9 Exchange-Coupled Nd z2 Fe'4
Fe 14 B Based Films
Aylesworth et al. (1988,
C1988, 1989) investigated the magnetic and structural
properties of Nd2 (Fe,
CFe, Co) 14 B/Fe multi-layers prepared in a multi-gun
sputtering system. The substrates used for this study were mica and Ta, and
the individual layer thickness in the multi-layers were 50 - 200 A for the Nd 17
l7
CFe09 Coo 1)76 B7 and 50 - 200 A for the Fe or Ag with the total multiplayer
(Fe09
thickness of 1 IJm. The coercivity of the multi-layers decreased with increasing
Nanostructured Exchange-Coupled Magnets 249

Fe layer thickness. Parhofer et al. (1996) performed a study on Nd-Fe-B/Fe/


Nd-Fe-B tri-Iayers with individual layer thickness between 10 and 100 nm,
prepared by sputter deposition. They observed a remanence enhancement due
to exchange coupling between the' the- soft and hard magnetic layers as a function
<
of the Fe-interlayer thickness d. For d ~ 30 nm, a Fe-layer between two
hard magnetic layers of comparable thickness is completely coupled to the
hard magnetic phase.
Shindo et al. a!. (1996, 1997)
1997> fabricated multilayer Nd-Fe-B/a-Fe
Nd-Fe-B/ex-Fe films of
the form Ti (30 nm)/Fe (dFe)/[Nd-Fe-B
TiC30 (dNd-Fe-B )/Fe( d Fe ) ] (5/Ti
(dFe)/[Nd-Fe-B(dNd-Fe-B)/Fe(dFe)] (5/TiC30 00 nm)/
glass by means of radio frequency (rf) sputtering. The thickness d Fe and
dNd-Fe-B varied from 0 to 50 nm and from 0 to 100 nm, respectively.
Magnetization measurements revealed that the minor loops of the multilayer
films are reversible in a certain range of reverse fields except for films without
a-iron and films where dNd-Fe-B < 10 nm,
ex-iron nm. A micromagnetism calculation of
magnetization curve using Landau-Lifshitz-Gilbert
Landau-lifshitz-Gilbert equation has reproduced the
observed hysteresis loops including the exchange-spring behavior and the
dependence of the coercive field M He on d Fe and dNd-Fe-B' dNQ.Fe-B' The Nd-Fe-B/ex-Fe
Nd-Fe-B/a-Fe
bilayer structure divided into lOx lOx N z (N z = = 1- 12 depending on the layer
thickness) cells has been considered. The size of each cell was taken to be
equal to that of a crystal grain for x, y directions (within the film plane) and
5 nm for z direction (normal to the film plane). They consider the
magnetocrystalline
magnetocrystall ine anisotropy energy and Zeeman energy for each cell. The
magnetic easy direction of each Nd-Fe-B grain is assumed to be randomly
oriented, while the grains are coupled with each other through interfaces with
exchange interactions. Three kinds of exchange interactions between the cells
have been considered, in which two intralayer coupling J h (between
coupl ing constants Jh(between
cells of Nd-Fe-B) and JJ ss (of Fe) are fixed to be 2.0 2. 0 x 10- 2 J/ m2Z .. The
10 - z J/m
parameters of exchange constant J hS hs (between Fe and Nd-Fe-B cells) and

crystal grain size were adjusted so as to best reproduce the experimental


data. The interlayer coupl ing constant was taken to be J hs hs = O. 07 x J h' h. These
results suggest that the interlayer exchange-coupling strength (corresponding
to J hShS )) is about an order of magnitude smaller than that of the intralayer
couplings (corresponding to J h h or J ss )'
)' From the comparison between
experiment and calculation, they inferred that the interlayer exchange-coupling
strength is about 10% of the intralayer coupl ings, which are found to be almost
independent of d Fe and dNd-Fe-B'
Yang and Kim (1999) investigated the sandwich structures of Nd2 zFe14 B/
Fe/NdzFe14B films by means of a KrF excimer laser (A =
Fe/Nd2Fe14B = 248 nm) ablation
technique. The magnetic properties were measured as a function of the
thickness of the hard (Nd2 z Fe14 B) and soft (Fe) magnetic layers and of the
volume fraction. In the (x nm) [NdFeB]/(y nm) [Fe]/(x nm) [NdFeB]/
(1 0 0) Si structure, the thickness x was varied from 3. 6 to 54 nm, and the
250 W. Liu et al.

thickness y from 15 nm to - 112 nm. At y = = 15 - 20 nm, the volume fraction


of Fe is 61%
61 % -75
-75%, %, the sandwich structure exhibited an enhancement of M,/
M s and MH c as the result of the exchange coupling between the magnetic
layers. The mossbauer
mbssbauer investigation indicated that the top Nd-Fe-B layer was
characterized by a magnetic texture depending on both the thickness of the
layers and the crystallization of the ex-Fe layer in Nd-Fe-B/Fe/Nd-Fe-B tri-
layers (Steyaert et al. , 1999).
Because of the higher saturation magnetization of Fe65
Fe6S C0 35 , as compared
to pure Co and Fe, the alloy was also chosen as the soft magnetic layer
component of the multilayer films. Liu et al. (2002a) investigated the
magnetic properties of nanocomposite multilayer magnets of the Ti (20 nm)/
[NdDyFeCoNbB (15nm) M] x 20/Ti(20 nm)/(Si substrate) multilayer (M=
(15 nm) M]X20/Ti(20
Co, Fe65 C0 35 ) prepared by sputtering and subsequent heat-treatments. X-ray
diffraction reveals a random orientation of the Nd2 Fe14 B-type phase in almost
all the films. Different thicknesses x for Co layers and y for Fe65 C0 35 3S layers
were adopted in the multi-layers. High remanence is achieved in the
nanocomposite multilayer films consisting of the Nd2 Fe14 B-type phase and soft
magnetic phase for Co with 6 ~ x ~ 4 nm and for Fe65Co35 Fe6SC03S with 10 ~
y ~ 6 nm. The room-temperature hysteresis loops for the multi layers with the
form Ti(20 nm)/
nm) / [NdDyFeCoNbB(
[NdDyFeCoNbB(15 15 nm)Fe6SC03S(ynm)]
nm) Fe65 C0 35 (ynm)] xX20/Ti(20
20/Ti (20 nm)/(Si
nm) / (Si
substrate) annealedat625'C
annealed at 625'C for 1 min. are shown inFig.6.40.
in Fig.6.40. For Fe65Co35
Fe6SC03S
layer thickness of 2 nm, the magnetic behavior is similar to that of single
layer, because the hard phase is dominant. However, the remanence is higher
than that of single layer.
Ti(20 nm) I/ [A( I5 20/I Ti(20 nm)/(Si substrate)
15 nm)FeCo(y nm)]* 20
1000
- - Y=IO
--- y=IO
800 -+-Y=8
- - Y=8
- 0 - Y=6
600
~
.........
- . - Y=2
Uu
~ 400
:::
:::l
E
~ 200
c:
<:
0
.~ 0
.!::l
.~
N
0)
<: -200
c:
OJ)
co
2'"
:2 -400

-600
-800Eel;;~,.h;P"'::~
-800~~~;OOO::~
-1000 L-~_~_~_~_~~_~_~
'---~_~_~_~_~~_~_~
-20 -15 -10 -5 0 5 10 15 20
Applied field (kOe)

Figure 6.40 Hysteresis loops at room temperature for the thin films of Ti-buffered
NdDyFeCoNbB/Fe65 C035 multi-layers on the Si substrate annealed at 625
625'C
'C for 1 min.
Magnetic measurements are in the film plane (Liu
CLiu et al. , 2002a).
Nanostructu red ExchangeCoupled
Nanostructured Exchange-Coupled Magnets 251

The saturation magnetization and remanence enhance with increasing


thickness of the Fe65 C0 35 layers, and the remanence reaches 88.2 Am 2 /kg for
y = 10. The remanence of the film with y = 6 is slightly higher than that of the
Co-containing series mentioned above due to the higher magnetization of
Fe65 C0 35 compared to Co. In contrast, the coercivity of the former is only
0.57 T that is remarkably lower than the latter. TEM images show that the
average grain size of the hard phase is between 10 and 15 nm, but the
average grain size of the soft phase is larger than that of the hard magnetic
phase. The ratio M riM r / M s5 for all the samples is higher than O. 65. The
enhancement of the magnetic properties in the nanocomposite multilayer films
is attributed to the exchange coupling between the magnetically soft and hard
phases.
Meanwhile, LiuL iu et al. (2002b) also investigated the structure and
properties of nanocomposite multilayer magnets of the NdDyFeCoNbB : Fe
prepared by sputtering and subsequent heat-treatments. Figure 6.41 shows a
TEM bright-field image of the as-deposited Ti ( 1Onm ) I[ NdDyFeCoNbB
10nm )/[
(16nm)Fe(6nm)]X20/Ti(10nm)/
(16nm)Fe(6nm)]x20/Ti(10nm)1 (glass ceramics) multilayer thin film, in
which a high resolution TEM image is inserted at the lower left corner. From
the cross-section view, the distribution of the soft and hard phase layers is
very uniform. Selected-area electron diffraction analyses reveal that the wide
NdDyFeCoNbB layer is amorphous and the narrow one corresponds to Fe.
Clearly, the ratio and distribution of soft and hard phases can be well
controlled this way.

Figure 6. 41 TEM brightfield


bright-field image of the as-deposited Ti ( 10 nm) /[ I[ NdDyFeCoNbB
(16 nm)Fe(6 nm)] nm)/I (glass ceramics) multilayer thin film. The inset is a
20/Ti<100 nm)
nm) ] x 20/Ti(1
high resolution TEM image of this sample (Liu et al. , 2002b).

In NdDyFeCoNbB/ex-Fe
NdDyFeCoNbBIex-Fe multilayer on the glass ceramics substrates
575'C
annealed at 575 'c for 30 min, the XRD analysis shows a large amount of ex-Fe
due to the existence of Fe layers in the multilayer film. Almost all other XRD
252 W. Liu et al.

peaks of the films correspond to randomly oriented Nd2 Fe14 B-type phases.
Hysteresis loops at room temperature for Ti( 10 nm)/[NdDyFeCoNbB( 16 nm)
Fe( 4.5 nm) ] x 20 /Ti ( 10 nm) / (glass ceramics-substrate) and Ti ( 10 nm) /
[NdDyFeCoNbB (320 nm )J/Ti ( 10 nm )/( glass ceramics-substrate) films
annealed at 575 'c for 30 min are given in Fig. 6.42.6. 42. In comparison with the
result of single layer, the remanence enhancement is observed in the
multilayer films. The Ti( 10 nm) /[NdDyFeCoNbB( 16 nm)Fe( nm) Fe(4. 4.55 nm) ] x 20/Ti
(10 nm)/(glass-substrate) multilayer film show a high remanence of !JoM, =
/JoM, =
1.31
1. 31 T and a high maximum magnetic energy product of (BH) (BH)maxmax = 204 kJ/ m3 ,
kJ/m
which are much higher than those of the single layer [!Jo [/Jo M, = O. 85 T and
(BH)max = 117 kJ/m 33 ].
] . The coercivity of the former (/JoMH
(!JoMH ce =0. 77 T) is little
(!JOMM He = 1.
lower than that of the latter (/Jo 1. 1 T). The reduced remanence of the
multi layers is 0.78. It is clear that the exchange coupling between the layers
of soft and hard phases results in a significant enhancement of the remanence
and an increase of the maximum magnetic energy product in this type of the
nanocomposite magnets.

1.6

1.2

0.8

t:E
c:
<: 0.4
0
.~
.~
N
.
.~ 0.0
"0
0..
0-
-D.4
--D.4

-D.8
--D.8

-1.2

-1.6 L-a=~::--~

-2500-2000-1500-1000 -500 0 500 1000 1500 2000 2500


Applied field (kAhn)
(kAlin)

Figure 6.42 Room temperature hysteresis loops for Ti TI (10 nm) Ti/[ NdDyFeCoNbB
(16 nm) Fe(4.5
Fe( 4.5 nm) 20/Ti (10
nm)]] x 20/Ti( 10 nm) I (glass ceramics substrate) multilayer and Ti
nm)/(glass Ti((10
10
[NdDyFeCoNbB( 320 nm) ITi(
nmHil [NdDyFeCoNbB(320 ITi ( 10 nm) I (glass ceramics substrate) films annealed at
575C
575'C for 30 min (Liu et al. , 2002b).

To investigate the nanostructures of the multilayer magnets, a plan view


was also obtained from TEM examination. Electron diffraction has confirmed
that the hard magnetic phase has the N~ Fe14 B-type structure. Figures 6. 43a, b
show the plan view elemental maps acquired by electron energy-loss
Nanostructured Exchange-Coupled Magnets 253

spectroscopy (EELS). The intensity (brightness) corresponds to the


concentration of the elements concerned. Since only the hard phase contains
Nd, the Nd map in Fig. 6. 43a reveals the grain morphology of the hard phase
with a grain size about 40 nm. In contrast, the Fe map in Fig. 6. 43b reveals
ex-Fe phase could appear in
the morphology of the soft phase, indicating that a-Fe
the form of the continuous layer, in agreement with the results in Fig. 6. 41.41 .
Figure 6. 43c shows an MFM image of the same film. It shows a domain
structure with characteristic lengths of order 300 nm. It is concluded that
although the nanostructured soft and hard phases diffused into each other during
annealing, the exchange-coupled soft and hard composite nanostructure still
remains. The best magnetic properties are obtained for that annealed at 575,
while the remanence enhancement is observed for those annealed at both 575
'c. A relatively low annealing temperature results in imperfect exchange
and 550 C.
coupling, because the hard phase in the film is not formed completely. If the
annealing temperature is too high, the decoupled effect is clearly observed
because the continuous growth of a-Fe,
ex-Fe, which destroys the exchange coupling.

(a)

(b)

Figure 6.43 EELS map of Nd (a), Fe (b) and MFM image of (c) in Ti(10 nm)/
[NdDyFeCoNbS( 16 nm)
[NdDyFeCoNbB( nm)/Fe(4.
/Fe( 4.55 nm)
nm)]] x 20/Ti( 10 nm)/(Glass
nm) / (Glass ceramics substrate) multi-
layers annealed at 575'C for 30 min (Liu et al. , 2002b).

Therefore, the nanostructures of the multi layers can be well controlled by


preparing different layers and by adjusting the thickness of the layers for the
anneal ing at appropriate temperatures, the
soft and hard phases. After annealing
254 W. Liu et al.

mutually dispersed soft and hard phases can be formed in the partial layers of
ex-Fe phase appears also in the form of the
the multilayer magnets, while the a-Fe
continuous layer. The high maximum magnetic energy products achieved in this
work show much promise for applications of nanocomposite multi layers as
multilayers
permanent magnet thin films.
Briefly, in exchange coupled magnets, efforts should be directed at
optimizing the microstructure with a reasonable ratio of the hard to soft phases
and obtaining a fine and uniform distribution of the grains with a small size in
coupled soft and hard phases. According to Skomski and Coey (1993) and the
above-mentioned experimental results, to obtain the higher energy product in
nanocomposite magnets, new processing techniques should be found to
develop bulk anisotropic exchange coupled magnets. Such a result might be
more easily achieved in thin-film systems with soft and hard phases.

6. 4 Conclusions

In this chapter, we have summarized the key features of exchange-coupled


permanent magnets. Exploiting exchange coupling is an important tool to
improve the remanence of isotropic magnets (remanence enhancement) and to
enhance the saturation magnetization of oriented magnets. However, the
improvement of the remanence is limited by the requirement of maintaining a
minimum coercivity of the order of M s /2. The idea behind these systems is to
break out of the straitjacket of natural crystal structures by artificially
structuring new materials. The concept is similar to that of the 4f-3d
intermetall ics themselves, but on a different scale, where the atoms are
intermetallics
replaced by a mesoscopically structured hard-magnetic skeleton with surplus
anisotropy and small soft-magnetic blocks.
Analytical calculations yield a well-defined and realistic upper limit of the
order 1 MJ/m 3 to the energy product of permanent magnets. This energy
product is predicted for suitable multilayered and random hard-soft
nanostructures. As a rule, high energy products require well-textured and fine-
grained structures with a high fraction of the soft-magnetic high-magnetization
phase. These requirements are fairly well satisfied in recently developed high-
energy-product Fe-Pt films, where a FePt-based phase provides the
Fe3 Pt-based phase enhances the magnetization.
anisotropy and a soft Fe3Pt-based
However, thin-film magnets are comparatively cumbersome to produce and
difficult to turn into magnets of a desired size and shape. Furthermore, the use
of high-magnetization phases such as a-Fe ex-Fe and Fe65 C0 35 has remained a
challenge.
The main problem is to ensure a sufficiently high coercivity. The above-
mentioned energy-product predictions are val id for ideal nanostructures, but in
Nanostructured Exchange-Coupled Magnets 255

reality there are always imperfections which reduce the coercivity and
undermine the energy product. From a basic point of view, this coercivity
reduction is well understood, but in practice it is often difficult to identify the
structural features governing the coercivity of a given material. This task
amounts to the determination of the micromagnetic spin configurations as a
function of the magnets real structure. One issue is the exchange at grain
boundaries, which affects the coupling between nanograins and, indirectly,
the coercivity and the maximum energy product. Both continuum and layer-
resolved analytic calculations yield a quasi-discontinuity of the magnetization
between misaligned and incompletely exchange-coupled grains. This
discontinuity is accompanied by a moderate reduction of the intergranular
exchange and facilitates the switching of soft phase. By contrast, anisotropy
changes in the grain-boundary region have no major effect on the spin
structure, because the effect of anisotropy inhomogenities averages over at
least a few nm. Future developments in the field will exploit our ever-
increasing knowledge of the microstructure and reversal mechanisms of the
materials to fully realize the potential of magnetic nanostructures and to
explore new applications.

References
Alben R. , J. J. Becker and M. C. Chi. J. Appl. Phys. 49: 1653 (1978)
AI-Omari, I. A. and D. J. Sellmyer. Phys. Rev. B52: 3441 (1995)
Arai, S., T. Shibata, N. Koshizuka and M. Nagakura. IEEE Trans. Magn.
MAG-23:2299
MAG-23: 2299 (1987).
Anderson, P. W. Phys. Rev. B 109: 1492 (1958)
Araki, T. , T. Nakanishi and T. Umemura. J. Appl. Phys. 85: 4877 (1999)
Aylesworth, K. D. , Z. R. Zhao, D. J. Sellmyer and G. C. Hadjipanayis. J.
Appl. Phys. 64: 5742 (1988)
Aylesworth, K. D.,
D. , Z. R. Zhao, D. J. Sellmyer and G. C. Hadjipanayis. J.
Magn. Magn. Mater. 82: 48 (1989)
Bauer, J. , M. Seeger, A. Zern and H. KronmOlier.
KronmLiller. J. Appl. Phys. 80: 1667
( 1996)
Bernardi, J., G. F. Soto, J. Fidler, S. David and D. Givord. J. Appl.
Phys. 85: 5905 (1999)
Betancourt, J. I. and H. A. Davies. J. Appl. Phys. 85: 5911 (1999)
Billoni, O. V. , S. E. Urreta, L. M. Fabietti and H. R. Bertorello. J. Magn.
Magn. Mater. 187: 371 (1998)
Branagan,D. J. ,M. J. Kramer, Y. Tang, R. W. McCallum. J. Appl. Phys. 87:
6737 (2000)
Boller, H. and H. bsterreicher.
Osterreicher. J. Less-Common. Met. 103: L5 (1984)
Bouchet, G. ,J. Laforest, R. Lemaire, and J. Schweizer. C. R. Acad. Sci.
Paris 262: 1 (1966)
Bowden G. J. , J. M. L. Beaujour, S. Gordeev, P. A. J. de Groot, B. D.
256 W. Liu et al.

Rainford and M. Sawicki. J. Phys Condens. Matter. 12: 9335 (2000)


Bruck, E. , Q. F. Xiao, P. D. Thang, M. J. Tooner, F. R. de Boer, B. H. J.
Buschow. Physica B300: 215 (2001)
Buschow, B. H. J., W. Luiten, P. A. Naastepad and F. F. Westendrop.
Philips Tech. Rev. 29: 336 (1968)
Buschow, B. H. J. , W. Luiten, P. A. Naastepad and F. F. Westendrop. J.
Appl. Phys. 40: 4029 (1969)
Buschow, K. H. J. Phys. Status Sol idi A 7: 199 (1971)
Solidi
Buschow, K. H. J. In: K. H. J. Buschow. ed. Handbook of Magnetic
Materials Elsevier, North-Holland, Amsterdam, 463 ( 1997>
1997)
Burzo, E. Rep. Prog. Phys. 61: 1099 (1998)
Cabrera, G. G. and R. V. Falicov. Phys. Status Solidi B 61: 539 (1974)
Cadieu, F. J. , T. D. Cheung and L. Wickramasekara. J. Appl. App!. Phys. 57:
4164 (1985)
Cadieu, F. J., T. D. Cheung and L. Wickramasekara. J. Magn. Magn.
Mater. 54 - 57: 535 (1986a)
Cadieu, F. J.,
J. , T. D. Cheung, L. Wickramasekara and N. Kamprath. IEEE
Trans. Magn. 22: 752 (1986b)
Appl. Phys. 61: 4105 (1987)
Cadieu, F. J. J. App!. (1987>
Cadieu, F. J.,
J. , R. Rani, X. R. Qian and L. Chen. J. App!.
Appl. Phys. 83: 6247
(1998)
Callen E., Y. J. Liu and J. R. Cullen. Phys. Rev. B 16: 263 (1977)
Cebollada, A. , D. Weller, J. Sticht, G. R. Harp, R. F. C. Farrow, R. F.
Marks, R. Savoy and J. C. Scott. Phys. Rev. B 50: 3419 (1994)
Chang, W. C. and D. M. Hsing. J. Appl. Phys. 79: 4843 (1996)
Chang, W. C., D. Y. Chiou, S. H. Wu, B. M. Ma and C. O. Bounds.
Appl. Phys. Lett. 72: 121 (1998)
App!.
Charles, R. J. , D. L. Martin, L. Valentine and R. E. Cech. AlP Conf. Proc.
5: 1072 (1971)
Chen, Z. M., X. Meng-Burany and G. C. Haj ipanayis. Appl. Phys. Lett.
Hajipanayis.
75: 3165 (1999a).
Chen, Q. , B. M. Ma, B. Lu, M. Q. Huang and D. E. Laughlin. J. Appl.
Phys. 85: 5917 (1999b)
Chen, Z. M., Y. Zhang, G. C. Hajipanayis, Q. Chen and B. M. Ma. J.
Magn. Magn. Mater. 206: 8 (1999c)
Chen, Z. M., Y. Zhang and G. C. Hajipanayis. J. Magn. Magn. Mater.
219: 178 (2000a)
Chen, Z. M., Y. Zhang and G. C. Hajipanayis. J. App!. Appl. Phys. 88: 1547
(2000b)
Chen, Z. M., H. Okumura, G. C. Hajipanayis and Q. Chen. J. Alloys and
Comp. 327: 201 (2001
Compo a)
(2001a)
Chen, Z. M., H. Okumura, G. C. Hajipanayis
Haj ipanayis and Q. Chen. Appl. Phys.
Lett. 89: 2299 (2001b)
Cheng, Z. H., M. X. Mao, J. J. Sun, B. G. Shen, F. W. Wang, C. L.
Nanostructured Exchange-Coupled Magnets 257

Yang, Y. D. Zhang and F. S. Li. J. Phys.: Condens. Matter 7: 2303


(1995a)
Cheng, Z. H. , B. G. Shen, M. X. Mao, J. J. Sun, Y. D. Zhang and F. S.
Li. Phys. Rev. B52: 9427 ,( ( 1995b)
Cheng, Z. H., B. G. Shen, F. W. Wang, L. Cao, J. G. Zhao, W. S.
Zhan, M. X. Mao, J. J. Sun, F. S. Li and Y. D. Zhang. Phys. Rev.
B 51: 12,433 (1995c)
Chiriac, H. , M. Marinescu, K. H. J. Buschow, F. R. de Boer and E. Bruck.
J. Magn. Magn. Mater. 202: 22 (1999)
Chopra, K. L., M. R. Randlett and R. J. Duff. Phil. Mag. 16: 261 (1967)
Chudnovsky E. M. , W. M. Saslow and R. A. Serota. Phys. Rev. B 33: 251
(1986)
Coehoorn, R., D. B. De Mooij, J. P. W. B. Duchateau and K. H. J.
Buschow. J. Physique 49: C8-669 (1988)
Coehoorn, R. , D. B. De Mooij and C. De Waard. J. Magn. Magn. Mater.
80: 101 (1989)
Coene, W. , F. Hakkens, R. Coehoorn, D. B. De Mooij, C. De Ward, J.
Fidler and R. Coehoorn R. , D. B. de Mooij, J. P. W. B. Duchateau and
K. H. J. Buschow. J. de Physique 49: C-8 669 (1988)
Coey J. M. D. Rare-Earth Iron Permanent Magnets. University Press, Oxford
(1996)
Coffey, K. R., M. A. Parker and J. K. Howard. IEEE Trans. Magn. 31:
2737 (1995)
Craik, D. J. Platinum Met. Rev. 16: 129 (1972)
Crepo, P., V. Neu and L. Schultz. J. Phys. D: Appl. Phys. 30: 2298
((1997)
1997)
Croat, J. J. Appl. Phys. Lett. 39: 357 (1981)
Croat, J. J. J. Appl. Phys. 53: 3161 (1982)
Croat, J. J. , J. F. Herbst, R. W. Lee and F. E. Pinkerton. J. Appl. Phys.
55: 2078 (1984)
Cui, B. Z. , X. K. Sun, W. Liu, D. Y. Geng, Z. Q. Yang and Z. D. Zhang.
J. Alloys and Comp. 302: 281 (2000a)
Cui, B. Z. , X. K. Sun, W. Liu, Z. D. Zhang, D. Y. Geng and X. G. Zhao.
J. Phys. D: Appl. Phys. 33: 338 (2000b)
Cui, B. Z. , X. K. Sun, W. Liu, Z. D. Zhang, D. Y. Geng, X. G. Zhao, J.
P. Liu and D. J. Sellmyer. J. Appl. Phys. 87: 5335 (2000c)
Cui, B. Z. , X. K. Sun, L. Y. Xiong, W. Liu, Z. D. Zhang, Z. Q. Yang and
A. M. Wang. J. Mater. Res. 16: 709 (2001)
Dahlgren, M. , R. Grossinger,
Grbssinger, D. R. Cornejo and F. P. Missell. J. Appl.
Phys. 83: 6268 (1998)
Dahlgren, M. , R. Grbssinger,
Grossinger, E. de Morais, S. Gama, G. Mendoza, J. F.
L iu and H. A. Davies. IEEE Trans. Magn. 33: 3895 (1997)
Liu
Das, D. K. IEEE Trans. Magn. 5: 214 (1969)
Davies, H. A. , A. Manaf, M. Leonowicz, P. Z. Zhang, S. J. Dobson and
258 W. Liu et al.

R. A. Buckley. Nanostruct. Mater. 2: 197 (1993a)


Davies, H. A. , A. Manaf and P. Z. Zhang. J. Mater. Eng. and Performance
2: 579 (1993b)
Ding J., P. G. McCormick and R. Street. J. Magn. Magn. Mater. 124: L1
(1993a)
Ding, J., P. G. McComick and R. Street. J. Magn. Magn. Mater. 123:
L239 (1993b)
Ding, J. , P. G. McComick and R. Street. J. Alloys and Compo 191: 197
((1993c)
1993c)
Ding, J. , P. G. McComick and R. Street. J. Magn. Magn. Mater. 135: 200
(1994)
Fang, J. S. and T. S. Chin. IEEE Trans. Magn. 32: 4401 (1996)
Fang, J. S., M. S. Leu and T. S. Chin. J. Appl.App!. Phys. 83: 3731 (1998)
Fang, J. S., T. S. Chin, J. C. Shih, C. A. Chen and S. W. Yung. J.
Magn. Magn. Mater. 195: 588 (1999)
Farrow, R. F. C. , D. Weller, R. F. Marks and M. F. Toney, A. Cebollada
and G. R. Harp. J. Appl. Phys. 79: 5967 (1996a)
Farrow, R. F. C. , D. Weller, R. F. Marks and M. F. Toney. Appl. Phys.
Lett. 69: 1166 (1996b)
Fischer R. and H. Kronmliller.
Kronmuller. J. Magn. Magn. Mater. 184: 166 (1998)
Foner, S. , J. McNiff, Jr. D. L. Martin and M. G. Benz. Appl. Phys. Lett.
20: 447 (1972)
Franklin, A. D., A. E. Berkowitz and E. Klokholm. Phys. Rev. 94: 1423
(1954)
Franse, J. J. M. and R. J. Radwanski. In: K.H.J. Buschowed. Handbook
of Magnetic Materials. North-Holland, Amsterdam, Vol. 7, Chapter 5 p. 438
(1993)
( 1993)
Fukunaga H. and H. Inoue. Jpn. J. Appl. Phys. 31: 1347 (1992)
Fullerton, E. E., J. S. Jiang, M. Grimsditch, C. H. Sowers and S. D.
Bader. Phys. Rev. B58. 12,193 (1998a)
Fullerton, E. E., J. S. Jiang, C. H. Sowers, J. E. Pearson and S. D.
App!. Phys. Lett. 72: 380 (1998b)
Bader. Appl.
Fullerton, E. E. , J. S. Jiang and S. D. Bader. J. Magn. Magn. Mater. 200:
392 (1999)
Fuerst, C. D. , J. F. Herbst, C. B. Murphy and D. J. Van Wingerden. J.
Appl. Phys. 74: 4651 (1993)
App!.
Gabay, A.,
A. , Y. Zhang and G. C. Hadj
Hadjipanayis. App!. Phys. Lett. 76: 3786
ipanayis. Appl.
(2000)
Gabay, A. M., W. Tang, Y. Zhang and G. C. Hadjipanayis. Appl. Phys.
Lett. 78: 1595 (2001)
Gao, Y. H. , J. H. Zhu, Y. Q. Weng, E. B. Park and C. J. Yang. J. Appl.
Phys. 84: 4388 (1998a)
Gao, Y. H., J. H. Zhu, C. J. Yang and E. B. Park. J. Magn. Magn.
Mater. 186: 97 (1998b)
Nanostructured Exchange-Coupled Magnets 259

Gauder,D.
Gauder, D. R. ,M. H. Froning,R.
Froning, R. J. White and A. E. Ray. J. Appl. Phys. 63: 3522
(1988)
Girt, Er. , K. M. Krishnan, G. Thomas and Z. Altounian. Appl. Phys. Lett.
76: 1746 (2000)
Girt, Er. , K. M. Krishnan, G. Thomas, E. Girt and Z. Altounian. J. Magn.
Magn. Mater. 231: 219 (2001)
Givord, D. , H. S. Li, J. M. Moreau. Solid State Commun. 50: 497 (1984)
GolI, D., M. Seger and H. Kronmuller. J. Magn. Magn. Mater. 185: 49
(1998
(1998))
GolI, D., I. Kleinschroth, W. Sigle and H. Kronmuller. Appl. Phys. Lett.
76: 1054 (2000)
Goto, T. , Y. Ide, H. Abe, K. Watanabe, J. Onagawa, H. Yoshida and J.
M. Cadogan. J. Magn. Magn. Mater. 198 198-199:
-199: 486 (1999)
Grossinger, R. J. Magn. Magn. Mater. 96: 189 (1991)
Gu, B. X. , B. G. Shen and H. R. Zhai. J. Appl. Phys. 63: 819 (1991)
Hadjipanayis, G. C. , R. C. Hazelten, and K. R. Lawless. J. Appl. Phys.
55: 2073 (1984)
Hadjipanayis, G. C. ,L. Withanawasam, R. F. Krause. IEEE Trans. Magn. 31:
3596 (1995)
Hadjipanayis,
Hadj ipanayis, G. C. J. Magn. Magn. Mater. 200: 373 (1999)
Harland, C. L.,L. , H. A. Davies. J. Appl. Phys. 87: 6116 (2000)
Hayase, M. , M. shiga and Y. Nakamura. Status solidi B 46: Kl17 K 117 (1971)
Henkel, O. Phys. Stat. Sol. 7: 910 (1964)
Herbst, J. F. and J. J. Croat. J. Appl. Phys. 55: 3023 (1984)
Herbst, J. F. Rev. Mod. Phys. 63: 819 (1991)
Herget, C. Proc. 8 th Int. tnt. Workshop on Rare-Earth Magnets and Their
Applications. Dayton, OH, 407 (1985)
Hirosawa, S. , H. Kanekiyo and M. Uehara. J. Appl. Phys. 73: 6488 (1993)
Hong, M. H.,H. , K. Hono and M. Watanabe. J. Appl. Phys. 84: 4403 (1998)
Imry Y. and Sh.-K. Ma. Phys. Rev. Lett. 35: 1399 (1975)
Inoue, A. , A. Takeuchi, A. Makino and T. Masumoto. IEEE Trans. Magn.
31: 3626 (1995)
Sol ina, V. A. Demshina and L. M. Msgat. Phys. Met.
Ivanov, O. A. , L. V. Solina,
Metallog. 35: 81 (1973)
G. , B. X. Gu and Y. W. Du. J. Appl. Phys. 88: 7230 (2000)
Ji, Q. G.,
Jiang, H. , J. Evans, M. J. O'Shea and J. H. Du. J. Magn. Magn. Mater.
224: 233 (2001)
Jiang, J. S., E. E. Fullerton, M. Grimsditch, C. H. Sowers and S. D.
Bader. J. Appl. Phys. 83: 6238 (1998)
Jakubowicz, J. , M. Jurczyk, A. Handstein, D. Hinz, O. Gutfleisch and K.
H. MOiler.
Muller. J. Magn. Magn. Mater. 208: 163 (2000)
Jakubowicz, J. J. Alloys and Compo 314: 305 (2001)
Jurczyk, M. and J. Jakubowicz. J. Magn. Magn. Mater. 185: 66 (1998)
Jurczyk, M. J. Alloys and Comp.Compo 299: 283 (2000)
260 W. Liu et al.

Jurczyk, M. and J. Jakubowicz. J. Alloys and Comp.Compo 311: 292 (2000)


Kanek iyo, H., M. Uehara and S. Hirosawa.
Kanekiyo, Hi rosawa IEEE Trans. Magn. 29: 2863
( 1993)
Kanekiyo, H. and S. Hirosawa. J. J Appl. Phys. 83: 6265 (1998)
Kaneko, Y. andN. Ishihaki. J. Mater. Eng. Perform. 3: 228 (1994)
Kapitanov, B. A.,A. , N. V. Kornilov, Y. L. Linetsky and V. Y Y. Tsvetkov
Tsvetkov. J.
Magn. Magn. Mater. 127: 289 (1993)
Keavney, D. J., E. E. Fullerton, J. E. Pearson and S. S D. Bader.
Bader IEEE
Trans. Magn. 32: 4440 (1996)
Keavney, D. J. , E. E. Fullerton, J. E. Pearson and S. D. D Bader.
Bader J. Appl.
Phys. 81: 4441 (1997).
Kelly, P. E., K. O'Grady, P. I. Mayo and R. W. Chantrell Chantrell. IEEE Trans.
Magn. 25: 3881 (1989)
Kim, Y. B. , W. S. Park, H. T. Kim, Y. S. S Cho, C. S. Kim, M. J. Park
and T. K. Kim. J. Appl. Phys. 77: 4133 (1995)
Kneller, E. F. and R. Hawig. IEEE Trans. Magn. 27: 3588 (1991)
Knoch,K.
Knoch, K. G.,
G. , B. Reinsh and G. Petzow, Proc. 13th13 th tnt. Workshop on Rare-
Earth Magnets and Their Applications. p. 503 (1992)
Kobayashi, T., M. Yamasaki and M. Harmano. J. Appl. Phys. 87: 6579
(2000)
Kojima, A. , A. Makino and A. Inoue. J. Appl. Phys. 87: 6576 (2000)
Knoch, K. G.,G. , B. Reinsch and G. Petzow
Petzow. ZZ. Metallkd. 85: 5 (1994)
Konno, T. J., M. Uehara, S. Hirosawa, K. Sumiyama and K. Suzuki. J.
Alloys and Comp. 268: 278 (1998)
Koon, N. C. and B. N. Das. App!. Lett. 39: 840 (1981)
Appl. Phys. Lett
Koon, N. C. and B. N. Das. J. Appl.App!. Phys. 55: 2063 (1984)
Kou, X. C., M. Dahlgren and R. R Grossinger. App!. Phys. 81: 4428
Gr6ssinger. J. Appl.
((1997)
1997)
Leineweber, T. , H. KronmOller.
Kronmuller. J.
J Magn. Magn. Mater. 176: 145 (1997)
Lemke, H.,H. , C. Echer and G. Thomas. IEEE Trans. Magn. 32: 4404 (1996)
Lewis, L. H., W. M. Bian, Y. Zhu and D. O. Welch. JJ. Appl. Phys. 19:
351 (1996)
Lewis, L. H. , D.D.O.O. Welch and V. Panachanathan. J. Magn. Magn. Mater.
175: 275 (1997)
Lewis, L. H. and V. Panachanathan. J. Appl. Phys. 85: 4883 (1999)
Li, W. , L. Jiang, D. Wang, T. Sun and J. Zhu. J. Less-Common Met. 126:
95 (1986)
Li, Z. W.,
W. , X. Z. Zhou, A. H. Morrish and B. G. Shen. Hyperfine Interact.
72: 111 (1992)
L' Heritier P. , P. Chaudonet, R. Madar, A. Rouault, J. P. Senateur and R.
Fruchart. C. R. Acad. Sci. , Paris, (1984)
Liu, J. F., T. Chui, D. Dimitrov and G. C. Hadjipanayis. Appl. App!. Phys. Lett
Lett.
73: 3007 (1998c)
Liu, J. P., Y. Liu, C. P. Luo, Z. S. Shan and D. J. Sellmyer. J. Appl.
Nanostructured Exchange-Coupled Magnets 261

Phys. 81: 5644 (1997a)


Liu, J. P., Y. Liu, Z. S. Shan and D. J. Sellmyer. IEEE Trans. Magn. 33:
3709 (1997b)
Liu, J. P., C. P. Luo, Y. Liu and D. J. Sellmyer. Appl. Phys. Lett. 72:
483 (1998a)
Liu, J. P., Y. Liu and D. J. Sellmyer. J. Appl. Phys. 83: 6608 (1998b)
Liu, J. P., Y. Liu, R. Skomski and D. J. Sellmyer. J. Appl. Phys. 85:
4812 (1999a)
LiuJ. P., R.Skomski, Y. Liu and D.J.Sellmyer. J. Appl. Phys. 87: 6740
LiuandD.J.Selimyer.
(2000a)
Liu, S. , J. Yang, G. Doyle, G. Potts and G. E. Kuhl. J. Appl. Phys. 87:
6728 (2000b)
Liu, Y., B. W. Robertson, Z. S. Shan, S. Malhotra, M. J. Yu, S. K.
Renukunta, S. H. Liou and D. J. Sellmyer. IEEE Trans. Magn. 6: 4035
(1994)
( 1994)
Liu, Y. , R. A. Thomas, S. S. Malhotra, Z. S. Shan, S. H. Liou and D. J.
Sellmyer. J. Appl. Phys. 83: 6244 (1998d)
Liu, Y., J. P. Liu and D. J. Sellmyer. Nanostructured Materials 12: 1027
(1999b) .
Liu Y. N., M. P. Dallimore, P. G. McCormick and T. Alonso. J. Magn.
Magn. Mater. 116: L320 (1992)
Liu, W. , Z. D. Zhang, J. P. Liu, X. K. Sun, X. G. Zhao and B. Z. Cui. J.
Phys. D: Appl Phys. 32: 2846 (1999c)
Liu, W. , Z. D. Zhang, J. P. Liu, X. K. Sun, D. J. Sellmyer and X. G.
Zhao. 221: 278 (2000c)
Liu, W.,
W. , Z. D. Zhang, J. P. Liu, X. Z. Li, X. K. Sun and D. J. Sellmyer.
J. Appl. Phys. 91: 7890 (2002a)
Liu, W. , Z. D. Zhang, J. P. Liu, L. J. Chen, L. L. He, Y. Liu, X. K. Sun
and D. J. Sellmyer. Advanced Materials 14: 1832 (2002b)
Lu, B. , M. Q. Huang, Q. Chen, B. M. Ma and D. E. Laughlin. J. Magn.
Magn. Mater. 195: 611 (1999a)
Lu, B. , M. Q. Huang, Q. Chen, B. M. Ma and D. E. Laughlin. J. Appl.
Phys. 85: 5920 (1999b)
Luo, c.C. P. and D. J. Sellmyer. IEEE Trans. Magn. 31: 2764 (1995)
Luo, H. L. J. Less Common Metals. 15: 299 (1968)
Manaf, A.,
A. , M. AI-Khafaji, P. Z. Zhang, H. A. Davies, R. A. Buckley and
W. M. Rainforth. J. Magn. Magn. Mater. 128: 307 (1993b)
Manaf, A. , R. A. Buckley and H. A. Davies. J. Magn. Magn. Mater. 128:
302 (1993a)
Massalski, T. B., Binary Alloy Phase Diagrams, 2nd ed, Vol. 2, p. 1755
((1990a)
1990a)
Massalski, T. B., Binary Alloy Phase Diagrams, 2nd ed, Vol. 2, p. 1241
(1990b)
Massalski, T. B., Binary Alloy Phase Diagrams, 2nd ed, Vol. 2, p. 1224
262 W. Liu et al.

(1990c)
Mcintyre, P. C. , C. J. Maggiore and M. Nastasi. J. Appl. Phys. 77: 6201
(1995)
Mcintyre, P. C., C. J. Maggiore and M. Nastasi. Acta Mater. 45: 869
( 1997)
Melsheimer, A., M. Seeger and H. Kronmuller. Kronmi..iller. J. Magn. Magn. Mater.
202: 458 (1999)
Mendoza-Suarez, G. and H. A. Davies. J. Alloys and Comp. Compo 281: 17 (1998)
Mendoza-Suarez, G., H. A. Davies and J. I. Escalante-Garcia. J. Magn.
Magn. Mater. 218: 97 (2000)
Miao, W. F. , J. Ding, P. G. McCormick and R. Street. J. Appl. Phys. 79:
2079 (1996)
Miao, W. F., F. , J. Ding, P. G. McCormick and R. Street. J. Appl. Phys. 82:
4439 (1997)
Mishra, R. K., G. Thomas, T. Yoneyama, A. Fukuno and T. Ojima. J.
Phys. Appl. 52: 2517 (1981>(1981)
Mishra, R. K. and V. Panchanathan. J. Appl. Phys. 75: 6652 (1994)
Mitani, S., K. Takanashi, M. Sano, H. Fujimori, A. Osawa and H.
Nakajima. J. Magn. Magn. Mater. 148: 163 (1995)
Muller,
Mi..iller, K. -H. , J. Schneider, A. Handstein, D. Eckert, P. Nothnagel and H.
R. Kirchmayr. Mat. Sci. Eng. A 133: 151 (1991) (1991>
Murakami, R. K. and V. Villas-Boas. J. Appl. Phys. 87: 6582 (2000)
Neu, V. , U. Klement, R. Schafer, J. Eckert and L. Schultz. Mater. Lett.
26: 167 (1996)
Neu, V. and L. Schultz. J. Phys. Appl. 90: 1540 (2001> (2001)
Nieber, S. and H. Kronmuller.
Kronmi..iller. Phys. Stat. Sol. (b) 153: 367 (1989)
O'Grady, K. , M. EL-Hilo and R. W. Chantrell. IEEE Trans. Magn. 29: 2608
(1993)
O'Sullivan, J. F. , P. A. I. Smith, S. David, D. Givord and J. M. D. Coey.
196-197: 182 (1999)
Panagiotopoulos, I. , L. Withanawasam and G. C. Hadjipanayis. J. Magn.
Magn. Mater. 152: 353 (1996)
Parhofer, S. M., J. Wecker, C. Kuhrt, G. Gieres and L. Schultz. IEEE
Trans. Magn. 32: 4437 (1996)
Parhofer, S. M., M. , C. Kuhrt, J. Wecker, G. Gieres and L. Schultz. J. Appl.
Phys. 83: 2735 (1998)
Ping, D. H., K. Hono and S. Hirosawa. J. Appl. Phys. 83: 7769 (1998)
Ping, D. H., K. Hono, H. KanekiyoandS. Hirosawa. Acta Mater. ~7: 4641
(1999)
Popov, A. G. , A. V. Korolev and N. N. Shchegoleva. Phys. Metallography
(USSR) 69: 100 (1990)
Ray, A. E. and K. J. Strnat. IEEE Trans. Magn. 8: 516 (1972)
Ray, A. M., Y. Zhang and G. C. Hadjipanayis, Appl. Phys. Lett. 76: 3786
(2000)
Nanostructured Exchange-Coupled Magnets 263

Rellinghaus, B. , J. Kastner, T. Schneider, E. F. Wassermann and P. Mohn.


Phys. Rev. B 51: 2983 (1995)
Rudman, P. S. and B. L. Averbach. Acta Metallurgica 5: 65 (1957)
Rodewald, W. , B. Wall, M. Kafler, K. ustUner and S. Steinmetz. In: G. C.
Hadjipanayis and M. J. Bonder eds. Rare Earth Magnets and Their
Applications Rinton Press, Princeton, 25 (2002)
Sabiryanov, R. F. and S. S. Jaswal. J. Magn. Magn. Mater. 177-181:
177 -181: 989
( 1998a)
Sabiryanov, R. F. and S. S. Jaswal. Phys. Rev. B 58: 12,071 (1998b)
Sagawa, M.,
M. , S. Fujimura, M. Togawa, H. Yamamoto and Y. Matsuura. J.
Appl. Phys. 55: 2083 (1984)
Sagawa, M. and S. Hirosawa, In: S. G. Sankar, J. F. Herbst and N. C.
Koon eds. High Performance Permanent Magnet Materials, Materials
Research Society Symposia Proceedings Materials Research Society,
Pittsburgh, Vol. 96, p. 161 (1987)
Sakuma, A. J. Phys. Soc. Japan. 63: 3053 (1994)
Sawicki, M. , G. J. Bowden, P. A. J. de Groot, B. D. Rainford, J. M. L.
Beaujour, R. C. C. Ward and M. R. Wells. Phys. Rev. B 62: 5817 (2000)
Schaller, H. J.,
J. , R. S. Craig and W. E. Wallace. J. Appl. Phys. 43: 3161
(1972)
Schneider, G, G. Martinek, H. H. Stadelmaier and G. Petzow, In: C.
Herget, H. Kronmuller and R. Poerschke eds. Proc. 6 th Int. Symp. on
Magnetic Anisotropy and Coercivity in Rare-Earth Transition Metal Alloys.
Bad soden p. 347 (1987)
Schrefl, T. , J. Fidler and H. Kronmuller. Phys. Rev. B 49: 6100 (1994)
Schrefl, T. and J. Fidler. J. Magn. Magn. Mater. 177-181: 970 (1998)
Schultz, L.,
L. , J. Wecker and E. Hellstern. J. Appl. Phys. 61: 3583 (1987)
Sellmyer, D. J. and S. Nafis. Phys. Rev. Lett. 57: 1173 (1986)
Sellmyer, D. J., M. Yu, R. A. Thomas, Y. Liu and R. D. Kirby. Phys.
Low-Dim. Struct. 1-2: 155 (1998)
Sellmyer, D. J. , M. Zheng and R. Skomski.
Skomsk i. J. Phys. : Condens. Mater. 13:
433 (2001)
Shen, B. G., L. Y. Yang, J. X. Zhang, B. X. Gu, T. S. Ning, F. Wo, J. G.
Zhao, H. Q. Guo and W. S. Zhan. Solid state Communications 74: 893
(1990)
Shindo, M. , M. Ishizone, H. Kato, T. Miyazaki and A. Sakuma. J. Magn.
Magn. Mater. 161: L1 (1996)
Shindo, M. , M. Ishizone, A. Sakuma, H. Kato and T. Miyazaki. J. Appl.
Phys. 81: 4444 (1997)
Shoemaker, C. B., D. P. Shoemaker and R. Fruchart. Acta Crystallogr.
Sect. C 40: 1665 (1984)
Skomski, R. Phys. Stat. Sol. (b) 174: K77 (1992)
Skomski R. and J. M. D. Coey. Phys. Rev. B 48: 15,812 (1993)
Skomski, R. J. Appl. Phys. 76: 7059 (1994)
264 W. Liu et al.

Skomski, R. J. Magn. Magn. Mater. 157-158: 713 (1996)


Skomski, R., J. P. Liu, J. M. Meldrim and D. J. Sellmyer. In:l.In:L. Schultz
MOiler eds. Magnetic Anisotropy and Coercivity in Rare-Earth
and K. -H. Muller
Transition Metal Alloys. Werkstoffinformationsgesellschaft, Frankfurt/M.,
p. 277 (1998)
Skomski, R. J. Appl. Phys. 83: 6503 (1998)
Skomski, R., J. P. Liu and D. J. Sellmyer. Phys. Rev. B60: 7359 (1999)
Skomski, R. and J. M. D. Coey. Permanent Magnetism. Institute of
Physics, Bristol (1999)
Skomski, R. and D. J. Sellmyer. J. Appl. Phys. 87: 4756 (2000)
Skomski, R. In: M. Ziese and M. J. Thornton eds. Spin Electronics.
Spr inger, Berl in, p. 204 (2001)
Skomski, R., H. Zeng and D. J. Sellmyer. IEEE Trans. Magn. 37: 2549
(2001)
Smith, P. A. I., J. Ding, R. Street and P. G. McCormick. Scripta Metall.
34: 61 (1995).
Sort, J. , J. Nogues, S. Surinach, J. S. Munoz, M. D. Bar6, E. Chappel,
Lett. 79: 1142 (2001).
F. Dupont and G. Chouteau. Appl. Phys. lett.
Steyaert, S., J. M. le Le Breton, S. Parhofer, C. Kuhrt and J. Teillet,
Tei lIet, J.
196 -197: 48 (1999)
Magn. Magn. Mater. 196-197:
Strnat, K. J. Cobalt (Engl. Ed.) 36: 133 (1967)
Strnat, K. J. IEEE Trans. Magn. 8: 51 (1972)
Strnat, K. J. Rare Earth-Cobalt Permanent Magnets. North-Holland,
Amsterdam,Vol.
Amsterdam, Vol. 4, Chap.2,
Chap. 2, p. 154 (1988)
Sui, Y. C. , Z. D. Zhang, Q. F. Xiao, W. Liu, X. G. Zhao, T. Zhao and Y.
C. Chuang. J. Phys.:
Phys. : Condens. Mat. 9: 9985 (1997)
Sumiyama, K.,K. , M. Shiga and Y. Nakamura. J. Magn. Magn. Mater. 31-
34: 111 (1983)
Sun, X. K.,
K. , J. Zhang, Y. l. L. Chu, W. Liu, B. Z. Cui and Z. D.D Zhang.
Lett. 74: 1740 (1999)
Appl. Phys. lett.
Tanaka, Y., N. Kimura, K. Hono, K. Yasuda and T. Sakurai. J. Magn.
Magn. Mater. 170: 289 (1997)
Tokunaga, M. , M. Tobise, N. Meguro and H. Harada. IEEE Trans. Magn.
22: 904 (1986)
Uehara, M., S. Hirosawa, H. Kanekiyo, N. Sano and T. Tomida.
Nanostructured Mater. 10: 151 (1998b)
Velu, E. M.,
M. , R. T. Obermyer, S. G. Sankar and W. E. Wallance. J. Less- less-
Common Met. 148: 67 (1989)
Uehara, M. , T. J. Konno, H. Kanekiyo, S. Hirosawa, K. Sumiyama and K.
Suzuki. J. Magn. Magn. Mater 177 177-181:
-181: 997 (1998a)
Villars, P. , A. Prince and H. Okamoto. Handbook of Ternary Alloy Phase
Diagrams. Vol. 5, p. 5612 (1995)
V. , S. A. Romero and F. P. Missell. J. Appl. Phys. 81: 4434
Villas-Boas, V.,
( 1997)
Nanostructured Exchange-Coupled Magnets 265

Visokay, M. and R. Sinclair. Appl. Phys. Lett. 66: 1692 (1995)


Wang, Z. C. , M. C. Zhang, F. B. Li, S. Z. Zhou, R. Wang and W. Gong.
J. Appl. Phys. 81: 5097 (1997)
Wang, Z. C.,
C. , S. Z. Zhou, M. C. Zhang, Y. Qiao and R. Wang. J. Appl.
Phys. 86: 7010 (1999)
Wang, Z. C., S. Z. Zhou, Y. Qiao, M. C. Zhang and R. Wang. J. Magn.
Magn. Mater. 218: 72 (2000a)
Wang, Z. C. , S. Z. Zhou, Y. Qiao, M. C. Zhang and R. Wang. J. Alloys
Compo 299: 258 (2000b)
and Comp.
Wang, Z. C., M. C. Zhang, S. Z. Zhou, and Y. Qiao and R. Wang. J.
Alloys and Compo
Comp. 309: 212 (2000c)
Wang, Z. C. , S. Z. Zhou, M. C. Zhang and Y. Qiao. J. Appl. Phys. 88:
591 (2000d)
Warren, B. E. X-Ray Diffraction. Dover Publications, New York, p.208 p. 208
(1990)
Watanabe, K. and H. Masumoto. Trans. Jpn. Inst. Met. 24: 627 (1983)
Watanabe, K. and H. Masumoto. Trans. Jpn. Inst. Met. 26: 362 (1985)
Watanabe, M., T. Nakayama, K. Watanabe, T. Hirayama and A.
Tonomura. Mater. Trans. JIM, 37: 489 (1996)
Watanabe, M. and M. Homma. Jpn. J. Appl. Phys. 35: L 1264 (1996)
Wecker, J. , M. Katter and L. Schultz. J. Appl. Phys. 69: 6058 (1991)
J. , K. Schnitzke, H. Cerva and W. Grogger. Appl. Phys. Lett.
Wecker, J.,
67: 563 (1995)
Withanawasam, L., L. , G. C. Hadjipanayis and R. F. Krause. J. Appl. Phys.
75: 6646 (1994)
Withanawasam, L., A. S. Murthy and G. C. Hadjipanayis. IEEE Trans.
Magn. 31: 3608 (1995)
Withanawasam, L. , I. Panagiotopoulos and G. C. Hadjipanayis. IEEE Trans.
Magn. 32: 4422 (1996a)
Withanawasam, L., I. Panagiotopoulos and G. C. Hadjipanayis. J. Appl.
Phys. 79: 4837 (1996b)
Wohlfarth, E. P. J. Appl. Phys. 29: 595 (1958)
Xiao, Q. F., T. Zhao, Z. D. Zhang, E. Bruck,
BrUck, K. H. J. Buschow and F. R.
de Boer. J. Magn. Magn. Mater. 223: 215 (2001a)
(2001 a)
Xiao, Q. F., BrUck, F. R. de Boer and K. H. J. Buschow.
F. , P. D. Thang, E. Bruck,
Appl. Phys. Lett. 78: 3672 (2001b)
Yang, C. J., and E. B. Park. J. Magn. Magn. Mater. 168: 278 (1997)
Yang, C. J. and S. W. Kim. J. Magn. Magn. Mater. 202: 311 (1999)
Yang, C. J.,
J. , E. B. Park, Y. S. Hwang and E. C. Kim. J. Magn. Magn.
Mater. 212: 168 (2000)
M. , T. S. Chin and S. K. Chen. J. Appl. Phys. 76: 7071 (1994)
Yao, J. M.,
You, C. Y., X. K. Sun, W. Liu, B. Z. Cui, X. G. Zhao and Z. D. Zhang.
J. Phys. D: Appl. Phys. 33: 926 (2000)
Yung, S. W., Y. H. Chang, T. J. Lin and M. P. Hung. J. Magn. Magn.
266 W. Liu et al.

Mater. 116: 411 (1992)


Zhang, J. , B. G. Shen, S. Y. Zhang and H. W. Zhang. J. Phys. D: Appl.
Phys. 33: 3161 (2000a)
Zhang, J. , S. Y. Zhang, H. W. Zhang and B. G. Shen. J. Appl. Phys. 89:
2857 (2001 a)
(2001a)
Zhang, J. X. , B. G. Shen.
Shen, L. Y. Yang and W. S. Zhan. Phys. Status Solidi
A122: 651 (1990)
Zhang, W. and A. Inoue. J. Appl. Phys. 87: 6122 (2000)
Zhang, W., M. Matsushita and A. Inoue. Mater. Trans. JIM, 41: 696
(2000b)
Zhang, W.,
W. , M. Matsushita and A. Inoue. J. App!.
Appl. Phys. 89: 492 (2001b)
Zhang, W. Y. , S. Y. Zhang, A. R. Van, H. W. Zhang and B. G. Shen. J.
Magn. Magn. Mater. 225: 389 (2001c)
Zhang, W. Y., A. R. Van, H. W. Zhang and B. G. Shen. J. Alloys and
Compo 315: 174 (2001d)
Zhang, Z. D. , W. Liu, X. K. Sun, X. G. Zhao, Q. F. Xiao, Y. C. Sui and
T. Zhao. J. Magn. Magn. Mater. 184: 101 (1998)
Zhao, T., Q. F. Xiao, Z. D. Zhang, M. Dahlgren, R. Grbssinger.
Grbssinger, K. H. J.
Buschow and F. R. de Boer. Appl. Phys. Lett. 75: 2298 (1999)
Zhou, J. , R. Skomski, C. Chen, G. C. Hadjipanayis and D. J. Sellmyer.
App!. Phys. Lett. 77: 1514 (2000)
Appl.
Ziese, M. and M. J. Thornton. (Eds.) Spin Electronics. Springer, Berlin
(2001)
Zijlstra, H. J. Appl. Phys. 41: 4881 (1970)

We are grateful to our colleagues including Jian Zhou.


Zhou, J. Ping Liu, George Hadjipanayis and
Sitaram Jaswal for helpful discussions and assistance. We thank DOE, AFOSR, ARO, NSF,
and CMRA for support of this work.
7 High-Field Investigations on Exchange Coupling
in R-Fe Intermetallics and Hard/Soft
Nanocomposite Magnets

Hiroaki Kato, Terunobu Miyazaki, and Mitsuhiro Motokawa

7. 1 Introduction

The performance of permanent magnets has greatly been improved by


introducing rare-earth elements (R) to their constituents (Herbst, 1991). It is
doubtless that the high coercivity of these magnets comes from the large
CCEF) acting
magnetic anisotropy originating from the crystalline electric field (CEF)
on R ions with large orbital angular momentum. Magnetization measurements
up to the high-field region where the hard-axis magnetization saturates are
indispensable in order to obtain the basic insight of the magnetic anisotropy.
Since the discovery of sintered Nd-Fe-B magnets (Sagawa et al. , 1984), we
have been investigating systematically high-field magnetization processes in a
series of Nd2 zFe14B-type compounds using mainly single crystal samples
(Nakagawa et al.,
aI., 1990). On the other hand, we developed a method of
analyzing these magnetization curves, which consists of a simpl simplified
ified
Hamiltonian taking exchange and crystal field at R ions into account, with Fe
sublattice being treated phenomenologically (Yamada et al., aI., 1988). Similar
analyses were made by Givord et al. (1988), Cadogan et al. (1988) and
Radwanski et al. (1990b). A comparison of crystalline magnetization process
(CEF) and molecular field parameters determined by these groups has been
made by Nakagawa et al. (1990). The essential feature in these analysis
models is the coupling of the two different types of sublattices. One is the R
sublattice,
sublattice , which gives large magnetic anisotropy owing to the CEF
interaction. Another is the Fe sublattice, which determines the large
magnetization and high Curie temperature as a result of strong Fe-Fe
exchange. The interactions between the two sublattices via the R-Fe exchange
coupling leads to a variety of magnetic properties such as spin reorientation
(SR) transitions and first-order magnetization process (FOMP). The SR is a
temperature-induced change in the easy-axis direction, observed, e. g. in an
RzFe14B
R2Fe14B system where R=Nd, Ho, Er and Tm (Hirosawa et aI., al. , 1986). The
origin of such SR transition is the competing magnetic anisotropy arising from
the second and higher order CEF potentials in the R sub
sublattice
lattice and from the Fe
268 Hiroaki Kato et al.

sublattice
sublattice,, which give different temperature dependences (Yamada et al. ,
1988). When the external magnetic fields are applied along the hard direction
at low temperatures, magnetization-jump behavior is observed at a certain
critical field value in some systems. Such a behavior is the FOMP (Asti and
Bolzoni, 1980), which is a result of discontinuous rotation of magnetic
moments and originates in a double-minimum behavior of higher order CEF
potentials (Yamada et al. , 1988).
It is well known that a large variety of solid solutions based on the
Nd2Fe14
Fel4 -B-type of structure has been investigated (Herbst, 1991). Our model
of analysis has proven to be applicable
appl icable not only to the R2M 14 B system where M
=Fe or Co (Kato et ai., al., 1988), but also to the CR1-xR'x)2FeI4B
CR 1 - x R'x)2 Fel4 B solid
solution system (Lim et al., 1991). Moreover, magnetic properties in an
interstitially-modified Sm2 Fell
Fel7 Nx system (Coey and Sun, 1990) have also
been explained by the same model calculations (Kato et al , 1993).
In general, such SR transitions will be accompanied by a considerable
lattice deformation, since there is a large orbital contribution to the R magnetic
moments, resulting in a strong coupling
coupl ing between spin and lattice systems. It is
therefore of interest to investigate the magnetoelastic properties (Clark,
1979) of these materials. In Section 7. 2, we present a general method of
simultaneously calculating magnetization and magnetostriction in a (R, R') -Fe-
X solid solution system, where X is a non-magnetic element such as B or N. In
Section 7. 3, experimental data on the high-field magnetization process and
the magnetostriction around the SR transition in a CErl-x CErl-X Tb x )2 Fel4
Fe14 B system
are shown and compared with the calculations based on the method given in
Section 7.2.
In Fig. 7. la, a schematic diagram of a high-performance mechanism in a
Nd2Fe14 B system is shown, displaying a cooperation, via J RFe , of different
types of sublattices with high anisotropy (R = Nd) and high magnetization
(Fe). In order to further improve a performance of permanent magnets without
finding a new compound beyond Nd2Fe14B, Fel 4B, this idea of the two-sub lattice-
coupl ing is extended to the artificial two-phase magnets. This is the exchange-
coupling
spring magnet, or nanocomposite magnet, proposed by Kneller and Hawig
( 1991 ). High-coercivity and high-magnetization permanent magnets can be
realized by combining the nanoscaled particles of "hard" phase, which has a
large coercivity (e. g. , Nd2 2 FeI4
Fe14 B)
B),, and"
and "soft"
soft" phase with high magnetization
(e. g. , a-Fe).
ex-Fe). Of course, these two phases must be magnetically" coupled"
magnetically "coupled"
via the inter-phase exchange interactions J hs ,' as shown in Fig.7.1b.Fig. 7. 1b. Many
works on the nanocomposite magnets have been reported especially after the
publication by Skomski and Coey (1993), in which they proposed a maximum
energy product beyond 100 MGOe in the completely aligned al igned ideal
nanocomposite system. Although an experimental realization of such an ideal
two-phase structure is still very difficult, it is worthwhile to introduce some of
the experimental efforts to fabricate the nanocomposite system and to compare
the micromagnetic calculations taking account of the inter-grain exchange
High-Field Investigations on Exchange Coupling. . . 269

7. 4, we describe the fabrication, magnetic properties


coupling. In Section 7.4,
and exchange coupling
coupl ing in multilayerd and nanodispersed Nd2Fe I4 B/a-Fe films.
Fe14

Anisotropy Magnetic moment


hard soft
(a)

Nanocomposite
-15nm

~~~ Coercivity Magnetization


(b)

7. 1 Schematic representation of the two-


Figure 7.1 two - sublattice coupling in (a)
Ca) Nd 2 Fe14 S,
Nd2Fe14 B,
and (b)
Cb) hard/ soft nanocomposite magnets. Characteristic length scales are about 1 nm
hard/soft
and 15 nm, respectively.

7. 2 Exchange and Crystal Field Model for


(R, R') -Fe-X System
(R,R')-Fe-X

In this section, we deal with the magnetic properties in a (Rl-xR:)hFekX


CR1-xR:)hFekX solid
solution system, which contains h (R CR1- x R' x) and k Fe atoms as magnetic
the case ofa R2Fe17N3 system, x=O, h=2, k=17, andX=N3 ,
elements. In thecaseofa
(Rl-xR'x)2FeI4B, h=4 and k=28, and X=B 2 , since the f and
whereas for CR1-xR'x)2FeI4B,
91 and
9 sites for the R or R' ions must be subdivided magnetically into ff,1 ,, f 2 , 9,
92 sites (Yamada
CYamada et al.,aI., 1988). The total free energy of the system at
temperature T in the external magnetic field H is assumed to be simply given
by
0.1)
(7.1)

In this equation, F RC R') is the free energy for the R (R')


F R')
R (F CR') sublattice written by
h

FR(H, T) =- k B T:Z:=ln:Z:=exp(-
FRCH, T2:ln2:expC- EsCi)/k B T) (7.2)
i= 1

EsC
in which k B is the Boltzmann constant and E j) is the 8-th
5 (i) s-th eigenvalue of the
following Hamiltonian for the i-th R ion,
270 Hiroaki Kato et al.

lIL
H R (;) =- i\L
HR(j) 5+ HCEF(i) + 21J
8+ HCEF(i) 21-1 s 5 H m (;) + I-Is(L
8 Hm(j) IJs(L + 25)
28) H
H
(7.3)

where, the first term expresses the spin-orbit interaction (L is the oribital
angular momentum, and 5 8 is the spin argular momentum) with the couplingcoupl ing
lI, while the second term HH CEF ( i; ) is the CEF Hamiltonian for the i-th
constant i\,
R ion written as

(7.4)
n.m

where A:;' (j) is the CEF coefficient and ~ V;;'


A;;' (i) V:;' (j) is the unnormalized tesseral
j

function for the j-th 4f electron. The non-zero terms in Eq. (7.4) (7. 4) are
determined depending on a point symmetry of the R or R' ion site and a choice
of quantization axes. The third term in Eq. (7.3) (7. 3) denotes the exchange
interaction between R and Fe spins, in which the molecular field for the i- th R
ion H m ( i) is proportional to the exchange integral J RFe and the Fe moment
m Fe . The matrix elements of H CEF ( i) and H R ( i) can be calculated using a
mFe'

tensor-operator technique as described in detail, e. g. by Yamada et al.


((1988).
1988). The third term in Eq. (7. 1) expresses the anisotropy and Zeeman
energy for the Fe sub lattice ,
sublattice,
(7.5)

where, K K Fe denotes the uniaxial anisotropy constant per Fe atom and e-the
angle between the z-axis and the direction of Fe moment mFe' The last term in
Eq. (7. 1) denotes an elastic energy for the volume change t::. /1 V / Va, in which
C is an elastic constant. As for the CEF Hamiltonian for the Rand R' sites, we
simply assume that CD only the dominant CEF coefficient M A~ is modified by the
volume change, and ~ the volume dependence of M A ~ (V) is determined by the
point charge approximation (PCA). In the PCA, A~ is inversely proportional
to the cube of a distance between 4f 4 f electron and a ligand
Iigand charge, which
directly leads to the following relation:
(7.6)

in which V V== Va + t::.


/1V. V. On the condition that the total free energy of Eq. (7. 1)
takes a minimum, we can simultaneously calculate t::. /1 V/ V ("" /1
t::. V/ Va ), and
Va),
the magnitude and direction of magnetic moments in each sublattice for various
temperature and field values.
In the case of a R2 Fe,
Fe! 48 system, systematic values of A:;' A;;' have been
obtained by comparing the high-field magnetization curves of single crystal
samples with the same calculation as presented here, without taking into
account the magnetoelastic interaction, i.e., C=O and t::.V=O /1V=O (Yamada et
al. , 1988). They found that the observed magnetization curves at 4.2 4. 2 K and
296 K including FOMP and SR for R = Pr, Nd, Sm, Tb, Dy, Ho, Er and Tm
compounds can be reproduced by using the same values of M" and A 4", if A;;' A:;'
High-Field Investigations on Exchange Coupling... 271

are normalized by the corresponding value estimated by PCA. The ratio of


final A~, A2'z
A;z and A~ to the PCA values are O. 15, 0.2 and 2.
2.0,
0, respectively,
which does not depend on R. In the case of the sixth order CEF coefficient A~
Ar ,
final values are 20 to 80 times larger than the PCA ones, in which the ratio
depends on R. Although it is well known that PCA is not directly appl
applicable
icable in
metallic system, the above results suggest that it is still a good measure, if
renormal ized properly, to evaluate a quantitative behavior and/or an R
renormalized
dependence in a certain series of R-Fe compounds, at least for lower order
CEF terms. In the derivation process of Eq. (7. 6), we thus adopt the PCA
only for the volume dependence of A~ (V),
( V), in which the magnitude of A~ itself
should be determined without PCA.

7. 3
7.3 High-Field Magnetization, Spin Reorientation
and Magnetostriction in (R,
(R,R')2Fe14B
R') 2Fe14B

7. 3. 1 Magnetic Phase Diagram and Spin Reorientation in


(Erl-x Tbx )2 Fe 14 B
Magnetic properties of RzFe14 B intermetallics have been extensively studied so
far, in connection with the high-performance Nd-Fe-B permanent magnets. In
the case of Erz Fe14 B, an abrupt SR transition at T SR = 323 K was observed
(Hirosawa and Sagawa, 1985), above which the direction of the magnetic
moments changes to the tetragonal [001] from [100]. Partial replacement of
Tb for Er is known to cause a rapid decrease of T SR and an appearance of an
intermediate phase with the tilted easy axis (Obermyer and Pourarian, 1991),
1991>,
as shown in Fig. 7.2. Such a behavior can be understood qualitatively by the
competing magnetic anisotropy arising from the Er, Tb and Fe sublattices
sublattices..
That is, the Er moment favors the easy [100] direction, owing to its positive
Stevens factor a J ' while that of Tb, with a J < 0, tends to align along the
[00 1]. The Fe moment, on the other hand, has been confirmed to favor the
[001].
[001]
[00 1] direction, accroding to the observed easy-axis behavior in the RzFe14 B
single crystals where R == Y, La, and Gd. In this section, we demonstrate how
the model calculation described in the preceding section can reproduce, or
sometimes predict, the experimental data in (Erl-x Tbx)zFe14B system.
(Erl-xTbx)2Fe14B
Polycrystalline samples of (Er,-x Tbx)zFe14B with a
(Erl-xTbx)zFe14B ~ x ='(
0 ='( ~ 0.6 were
prepared by melting under an argon atmosphere using an induction furnace.
Magnetically aligned samples were prepared by orienting the crushed powders
in a field of 10 kOe using epoxy resin. The temperature dependence of
magnetization was measured by using a vibrating-sample magnetometer with a
field of 1 kOe applied perpendicular to the aligned direction. High-field
magnetization measurements at 4.2 K were carried out by using a sample-
272 Hiroaki Kato et al.

400

ao M
,1.,//+,1.,1-
A;;+A1-

~200
- calc.

100

ao

o 0.3 0.4 0.5


x

Figure 7.2 Magnetic phase diagram of (Er'-xTbx)2Fe,48.


CErl-x Tb x >2Fe 14 B. Solid lines are results of
calculations.
calculations, while open and solid symbols are the experimental spin - reorientation
temperatures determined from magnetization and magnetostriction measurements. measurements,
respectively. i\A II and i\A1- are the longitudinal and transverse magnetostriction contants,
respectively.

extraction magnetometer in fields of up to 270 kOe, generated by a hybrid


magnet.
The solid lines in Fig. 7.2
7. 2 are phase boundaries calculated by using the
same procedure as described in the previous section, except that the volume
change 11 V/ V is not taken into account. The parameters used in this
calculation are given in Table 7. 1; which are identical with those reported
previously, (Yamada et al. , 1988). These calculations are in good agreement
with the experiments, although no extra parameters are assumed. Calculated
Fe'4 B with x = O. 4 are shown in
magnetization curves for (Er,-x Tb xx )2 Fe,4
Fig. 7. 3a, which exhibit abrupt increases in magnetization at 80 kOe, 130 kOe
and 290 kOe with fields applied along the [100J, [11 OJ and [001] directions,
respectively. Such FOMP' s are a result of discontinuous rotations of Er, Tb
and Fe moments caused by a double minimum behavior of CEF potentials
(Yamada et ai.,aI., 1988, Kato et ai., al., 1995). In order to confirm this
prediction, we measured the high-field magnetization curves. Figure 7. 3b
shows the experimental magnetization curves for the aligned polycrystal of
(Ero6Tbo4)2Fe14B.
(Ero6Tbo4)2Fe,4B. Although somewhat smeared, owing mainly to the
incomplete alignment, anomalous increase in magnetization was observed
around 70 and 240 kOe with field perpendicular and parallel to the aligned
that not only zero-
directions, respectively. Therefore, it has been confirmed thai
field properties such as SR, but also high-field magnetization behaviors in
these mixed compounds can be predicted successfully by using the simple-
mixture model. If we extend the field range, e. g. , up to 200 T, successive
phase transitions toward the forced-ferromagnetic alignment should occur as
shown in Fig. 7. 4. Similar successive phase transitions in high fields are
reported for Tm2 Fe'4
Fe14 B (Kato et aI.,
al. , 1995).
High-Field Investigations on Exchange Coupling. . . 273

7. 1
Table 7.1 CEF parameters A::' Ka 0- n) and molecular field H mm (K).
A ~ (in units of Kao (K) .
A,2
A,,2 A~ A~ A~
A: Hm
H

Er2 Fe14 B
Fe'4 B 283 -471 -12.8 - 0.980 -5.33
- 5. 33 145
Tb2Fe14 B 300 -462 -12.6 - O. 958 -5.17 145

(Er06 Tbo4 h Fel4 B


)2 Fe)4

';' 30~~~~L_--:::=
;- 30~""""':~~~ _ _-==:::t-_-
H//[OOI] _ _-=.....-_-==I_--
....:
..,;

~ 20
!:::' 20
:3
~
::s 10 calc.
OK
o 10 20 30
J-LoH
JloH (T)
(a)

H//Ha1ign
.-,
'""' 30
,,3 30 - /~-- - -----
:.H:/'.:.:'1H:.:a~li~gn~:::==- _ _~ -H.l-H
~
/
/--- H.l-Ha1ign
align

q,
ca, 20 //
.3 ./
~
~ //'H.l-H
/ / ' H.l-Halign
align
10 /I obs.
I
I 4.2 K
4.2K
I

o 10 20 30
JloH
J-LoH (T)
(b)

Figure 7. 3 (a) Calculated (T = 0 K) and (b) (T = 4. 2 K) observed magnetization


curves for (Er0.6
(ErO.6 Tb oo4 )2FeI4B.
. 4)2Fe 14 B.

120
Fel4 B
(Er06 Tbo4 )2 Fe)4 T=O K
T=OK
100
Fe
.-,
'""'
::i
....:
..,;
q,
ca,
80

60
Ed ,Tb
Ed tTb
I1 Fe

EJ~Tb
2>
.3
::s
~ 40

20

0 50 100 150 200


H (T)
J1f.1ooH(T)

7.44 Calculated magnetization curves at T = 0 K for (ErO.6 Tb o.4)2


Figure 7. Fe'4 B with field
) 2Fe14
parallel to [001]
COOl] direction up to 200 T.
274 Hiroaki Kato et al.

7. 3. 2
7.3.2 Magnetostriction and Spin Reorientation in
Tb x )2 Fe14B
(Erl-xTbx)2FeI4B
(Erl-x
Next we measured the magnetostriction in (Erl-x Tbx)2Fe14B system in order
to investigate the lattice deformation at the SR (Kato et ai., al., 1999.
1999, 2001a).
Ribbon samples of (Er (Er1- x Tb x) 2Fe14 B where 0 ~ x ~ O. 45 were prepared by
1- xTb

a single-roller rapidly quenching method in an Ar atmosphere. A surface


velocity of the copper wheel was varied between V s = = 4 - 31 m/s,
m/ s, ejection
pressure of Ar gas PAr = O. 6 - 1. 4 kg/ cm 2 . Magnetostriction measurements
were performed by a capacitance method apparatus (Ishio and Sato Sato,, 1988),
by which we observed the length change /::,./ t:./ / /, parallel (i\
(II II) and
(i\ 1-) to the field direction within the ribbon plane.
perpendicular (i\.l)
X-ray diffraction patterns of ribbons prepared by the conditions of V s =
15.7 m/s and PAr = 0.6 kg/cm 2 are shown in Fig. 7. 5. The (OO/)-type
= 0.6kg/cm (OOJ)-type
reflections are observed as major peaks, whi Ie small peaks of other types of
reflections such as (410) are additionally noticed. These results suggest that
the c-axis of the tetragonal cell aligns perpendicular to the ribbon surface.
Similar preferential crystallite orientation has been reported for Nd-Fe-B melt-
spun ribbons (Coehoorn and Duchateau, 1988). Since the ribbon thickness is
much larger than the X-ray penetration depth, the above data only provide the
information about the surface portion of the ribbons. In order to estimate the
bulk averages of the alignment, we measured the magnetization curves with
fields parallel and perpendicular to the ribbon surface. The full width at half
maximum of the c-axis orientation distribution is 12 5 and the degree of
alignment
al ignment was found to be not affected by the Tb replacement.

~
co
~
'0
0
0
0 80' S 2. 00-
~
00 80'
0

~
~
T
1 :!,
:!..
0
~
'{ x=o
0
2.
S

::i
:::i
~
.~
0
.;;;
cc: I A 0.15
ct:"
<J)

I 0.26
30 50 70
2en
Figure 7. 5 X-ray diffraction patterns for (Erl-,
CEr,-x Tb, )2Fe,4
),Fe14 B ribbons for samples with
x = O.
0, O. 15 and O. 26.

II
i\ II and II 1- are measured as a function of magnetic field at fixed
i\.l
High-Field Investigations on Exchange Coupling. . . 275

temperatures. Examples of the results for the x = 0 sample are shown in =


Fig. 7.6, in which the sum, i\i\11II + i\i-'
Fig.7.6, i\ 1- , and the difference, i\i\11 i\ 1- of the two
II - i\i-
constants are plotted against temperature. The value i\ II + i\ 1- i- corresponds to
the areal change of the ribbon plane, i. e. the c-plane, while i\ II - i\ i- 1- is an
anisotropic magnetostriction within the c-plane. From Fig. 7. 6 one can see
that overall values of i\ II + i\ 1- i- at 15 kOe are much larger than those at
2.5 kOe. This behavior is consistent with the report by Algarabel et al.
(1992)
e1992) and suggests the contribution from the volume magnetostriction owing
mainly to the Fe 3d band. The most significant feature in Fig. 7.6 is that there
exists a maximum value of i\ II + i\ 1- i- around 325 K. The temperature at which
this maximum value occurs slightly increases with increasing field. On the

x=0
10--(j
X 10--6

40
--i
<-<
~
+
~
20

0
280 320 360 400
T(K)

x=0
--i
<-<I
~
15
IS kOe 2.5 kOe
~
10

280
~I
0
~?
~?
000
0 0

320 360
:-::
::: 400
T(K)

(A II + ill.)
Figure 7. 6 Temperature dependence of the areal (All A1-) and anisotropic (ilil
(A II - ill.)
A1- )
ex
Er2 Fe14 B (x = 0) ribbon. The field H was applied within the ribbon
magnetostriction in Er2Fe14
plane. Solid lines are guides to the eye.

other hand.
hand, the anisotropic magnetostriction i\ II - i\ 1-
i- is almost constant in this
temperature and field range. Since the observed peak temperature of i\ II +
i- almost coincides to T SR of the x
i\ 1- =
0 sample, we can interpret the
phenomenon as the areal increase of the c-plane at the SR transition. Similar
=
plots for the x = O. 15 sample are shown in Fig. 7.7, in which i\ II + i\ 1-i- exhibits
a double peak at T, T 1"'"
-- 170 K and T 2"'"2 -- 100 K. Lim et al. (1992)
e1992) reported
eErl-x Tbx)2Fe14B with O. 05 ~ x ~ O. 15 in a
such successive transitions in (Erl-x
similar temperature range and assigned the higher and lower temperature
transitions as the SR's
SR' s from axial to tilted, and tilted to planar phases. The
observed double peak, therefore,
therefore. appears to correspond the areal increase of
rotation. that is, the two SR
the c-plane at the start and end of the easy-axis rotation,
transitions. Again, i\ II - i\ i-1- shows no anomaly at T 1 1 or T 2' but merely
276 Hiroaki Kato et al.

exhibits a broad peak around 150 K. Peak temperatures of i\ II + i\.l


i\ J including
other samples are plotted in Fig. 7. 2, which are in agreement with the
transition temperatures determined from magnetization data (Lim et al.,
1992) .

10-66
X 10-
20
-j
--j
~
+ 10
~
...<
0

80 120 160 200


T(K)

x 10-6 x=0.15
30
-j
--j
~
I 20
~
...< 2.5 kOe
10
0
80 120 160 200
T(K)
Figure 7.7 15
The same plot as that in Fig. 7.6 for CEr,- x Tb x )2 Fe" B with x = O. 15.

In order to analyze the experimental data above, we calculated the


volume magnetostriction arising from CEF potential of Er and Tb sites, as
described in Section 7. 2. The parameters given in Table 7. 1 are used as
before. As for the elastic constant C defined in Eq. (7. 1), we fixed it to be
2. 0 x 106 K by adopting the bulk modulus value reported for Er2Fe14B
2.0 Er2 Fe,4 B (Sidorov
and Khvostantsev, 1994). Figure 7.8 shows the calculated easy-axis direction
e and the volume change D. /1 V/ V as a function of temperature. For x = 0, as =
shown in Fig. 7. 8a, a discontinuous change of D. /1 V/ V is seen at T SR for H = O. =
With increasing the field strength, applied along the [100] direction, T SR
increases and the anomalous change in D. V// V around T SR becomes smaller.
/1 V
This result is in qualitative agreement with the experiments given in Fig. 7. 6a,
although observed i\ II + i\i\.l
~ does not show a discontinuous change but exhibits
just a maximum. In the case of x = = O. 15, calculated values of D./1 V/ V
(Fig. 7. 8b) exhibit a double peak at T, and T 2 . It should be noted that, with
increasing field, the peak at T 2 shifts to the higher values, whereas the T,
peak becomes smeared. This behavior around T2 is similar to the result for x = O.
Such calculated results appear to correspond to the observation (Fig. 7. 7a), 7a) ,
although the field dependence of observed T, peak is not obvious, owing to
the small number of observed data points. Calculated values of T T,1 and T 22 for
H = 0 are found to be almost equal to the solid lines in Fig. 7.2,
sol id Iines 7. 2, for which the
last term in Eq. (7. 1) was not taken into account. Thus, the effect of
magnetoelastic interaction on the magnitude of T SR has ~en confirmed to be
High.Field
High-Field Investigations on Exchange Coupling. . . 277

90
x=O
x=0 :""T--\
T""\"--\
~ \ \ 5 kOe
60 : \ "-
~
:. \2.5 kOe ............
\2.5kOe ...... -.
~ 30
H=O
..lkOe-,_
......lkOe-'_
--- --
........................
0

4
::..
:::..
s::<I<l 0 ::'rkO~/2-5kO
:.:h5-.
/ 2. . .-5kO e..-
_-
H=O e ....... - -
: (I / //5kOe
5 kOe
-4
-8X 10-6
10-n
L----- _....J.-_/
c...;...J--

280 320 360 400


T(K)
(a)

90
x=0.15
,-, 60
S
~~
30

:::.. -4
::..
s::<l<I
-8

-12X 10-6
IO-n
80 120 160
T(K)
(b)

Figure 7.8 Calculated temperature dependence of easy-axis direction 8 e and volume


change 1:..
b. V/ V in (Er,-x
(Erl- x Tb xx )2 Fe" = =
Fe14 B with (a) x = 0 and (b) x = O. 15. The field is applied
in the [100J direction (8(e = 90').
90) .

negligibly small in the present system.


It is suggested that 11 V// V
t::. V V exhibits a peak behavior at the SR transitions
to reduce the barrier height of the CEF potentials and to make the rotation of R
magnetic moments easier. In the present experiment, what we measured is
not a volume change 11 V// V, but an areal change within the c-plane. In order
t::. V
to assure the analysis and interpretation given above, information on the I1c t::.c / c
is also necessary. However, we consider the lattice deformation along the c-
axis to be less important, since the CEF potential at R site is dominated by the
neighboring R ions within the c-plane (Yamada et aI., 1988). In conclusion,
we have shown that magnetostriction constants i\A II + i\..L A -l take a maximum at
successive SR transitions, and these peaks can be explained as the volume
expansion owing to the CEF potential at R sites.
278 Hiroaki Kato et al.

7.
7.44 Exchange-Coupling in Hard/Soft Nanocomposite
Films

Despite the tremendous efforts to develop a high performance magnet beyond


a sintered Nd-Fe-B system, we are still on our way to finding such a new
system. A concept of exchange-spring magnets proposed by Kneller and
C1991) is a promising candidate to overcome such difficulty. Skomski
Hawig (1991)
and Coey (1993)
C1993) made an estimation of maximum energy product value in
exchange-coupled hard/soft nanocomposite magnets, which is more than twice
as large as that of the present Nd-Fe-B magnets. Experimentally,
nanocomposite magnets have been fabricated mainly by means of melt-
spinning or rapid-quenching process (Coehoorn a!. , 1988, 1989; Hirosawa
CCoehoorn et aI.,
et a!.,
al. , 1993; Withwanawasam et a!., al., 1994). At present, however, attained
maximum-energy-product values are still much lower than those expected
theoretically. The primary reason is thought to be a great gap between actual
and ideal nanostructures (good
Cgood alignment
al ignment of hard phases, suitable size of the
grains, and so on). This gap is caused by the difficulty in controlling
nanostructures in rapid-quenching process. In a thin-film process, it is
relatively easy to control nano-size structures. AI-Omari and Sellmyer (1995) C1995)
found the exchange-coupled behavior in CoSm/FeCo bilayer films, while a
remanence-enhancement behavior has been reported by Parhofer et al.,
(1996)
C1996) observed for Nd-Fe-B/Fe/Nd-Fe-B trilayers. Epitaxially grown hard/
soft bilayers and multi layers have been fabricated by Fullerton et al. (1998a,
C1998a,
1998b). We have also fabricated exchange-coupled hard/soft nanocomposite
films by RF magnetron sputtering, in which the c-axis of the hard grains are
both random (Shindo
CShindo et a!.
al.,, 1996, 1997; Ishizone et al., a!. , 1999) and oriented
CIshizone et al., 2000; Kato et al., 2000a, 2000b). In this section, we
(Ishizone
present two kinds of randomly-oriented Nd2FeI4B/ex-Fe
Nd2 Fe14 B/a-Fe nanocomposite films:
multi player geometry, while in the second
The first system has a hard/soft multiplayer
system the hard and soft grains are randomly dispersed in the film as in melt-
spun nanocomposite ribbons.

7.4.1 Multilayerd Nd2Fe14B/a-Fe Films

Multilayer films were fabricated by RF magnetron sputtering. The base


pressure of the sputtering system was below 4.0 x 4
X 10- Pa and the Ar pressure
1
during sputtering was 8.0 x 10- Pa. The films have the form of glass/Ti/[a-
glass/Ti/[ex-
Fe/Nd-Fe-BJ5 ex-Fe/Ti, with the thickness of both Ti layers d Ti = 30 nm, a-Fe
Fe/Nd-Fe-BJs / a-Fe/Ti, ex-Fe
= 0 - 50 nm, and Nd-Fe-B layers dNd-Fe-B ==
layers varied over the range of d Fe =
High-Field Investigations on Exchange Coupling. . . 279

o- 100 nm. Since as-sputtered Nd-Fe-B layers were amorphous, all the films
were annealed at T a = 873 K for 30 min with the pressure below 4.0 x 4
X 10- Pa
to crystallize the Nd2 Fe14 B phase.
According to the electron probe microanaysis (EPMA) CEPMA) experiments, the
composition of the as-sputtered single-layer Nd-Fe-B films was found to be
Nd 13
13 -- 15 Febal B7 - 11 , so the films are expected to be almost of single phase
containing Nd2 Fel4 Fe14 B grains. X-ray diffraction pattern for this film, shown in
Fig. 7. 9a, also exhibits a single-phase nature and random alignment of the Nd2
Fe14B phase. On the other hand, X-ray diffraction pattern for the multilayer
film exhibits an apparent two-phase structure of randomly oriented Nd2 Fe14 B
and a-Feex-Fe layers, as shown in Fig. 7. 9b. Shown in Fig. 7.10 are the dark-field
cross-sectional TEM images of [a-FeC [ex-FeC d Fe ) /Nd-Fe-BC30 nm) J5 multi layers with
d Fe = 5 and 30 nm. These images show the layered structure with well-defined
boundaries between Nd-Fe-B and a-Fe ex-Fe layers.

40 45 50 55 60
2e(")
28()
(b)

Figure 7.9 X-ray diffraction patterns obtained by using the Fe Ka


Kex radiation: (a) glass /
Ti (0.' ~m) / Nd-Fe-B (1.0
(0. llJm) ('.O~m)
IJm) / Ti (0. ~m) and (b) glass / Ti (30nm)
(0.'llJm) (30 nm) / [a-Fe
[ex-Fe
(50 nm) / Nd-Fe-B (100 nm) J5 / a-Fe
ex-Fe (50 nm) / Ti (30 nm).

Figures 7. lla11 a and 7. 12a show typical magnetization curves of the films of
[a-Fe
[ex-Fe (Cd Fe ) /Nd-Fe-B C30 nm) J5 multi layers where d Fe == 5 nm and 30 nm,
280 Hiroaki Kato et al.

L-...J
20 nm
20nm
(a)

L---...J
L---..J
50nm
(b)

Figure 7.10 TEM images of glass / Ti (30 C30 nm) / [ex-Fe (d


Cd Fe ) / Nd-Fe-B (30
C30 nm)]s/
nm)]5/
ex-Fe (d
Cd Fe (30 nm) multilayers: (a) d FFe.=5nm,
) / Ti (30nm)
F. ) =5 nm, and (b) d Fe =30nm.
=30 nm.

respectively. Although the major loop in the former case approximately


exhibits a form of a single hard phase, minor loop behaviors are completely
different. Namely, minor loops are reversible in a field range where the
magnetization is relatively large. On the other hand, the major loop of the
d Fe =30
dFe =-30 nm sample resembles a form of two independent magnetic phases with
reduced coercive field values. Minor loops are, however, almost reversible
even if the demagnetization field value exceeds the coercive field, and the
recoil permeability is quite high. This behavior is the one expected in
nanocomposite magnets in both small and large soft phase thickness,
respectively, was first discussed by Kneller and Hawig (1991). Thus, these
results strongly suggest that ex-Fe and Nd2 Fe14 B phases are exchange coupled
in these multilayer films. We have measured the magnetization curves of
various multilayer films with different d Fe and dNd-Fe-B
dFe dNcl-Fe-B systematically. Solid
circles in Fig. 7. 13 show the coercivity He as a function of d dFe
Fe and dNd-Fe-B
dNd-Fe-B for
the [ex-Fe ( d dFe
Fe ) /Nd-Fe-B ( dNd-Fe-B ) J5
dNcl-Fe-B ) J5 samples. Figure 7. 13a - c show
respectively the d Fe dependence with fixed dNd-Fe-B
dFe dNcl-Fe-B = 30 nm, the dNd-Fe.B
dNcl-Fe-B

dependence with fixed d Fe = 10 nm and the thickness d dependence when d =


dFe
dNd-Fe-B=d
dNcl-Fe-B = d Fe . As shown in Fig.7.13a,
Fig. 7. 13a, He decreases with increasing d Fe and
dFe
High-Field Investigations on Exchange Coupling...
Coupling. . . 281

the rate of decrease becomes significant for d Fe > 10 nm. In the case of Fig.
7. 13b, He is almost constant for dNd-Fe-B>
7.13b, dNd-F&-B> 20 nm, while, after taking a small
=
dNd-F&-B = 15 nm, it decreases drastically for dNd-Fe-B < 10 nm.
maximum around dNd-Fe-B
From Fig. 7. 13c, it is seen that He exhibits a maximum when the layer
thickness is about 10 - 20 nm. We have found that, for all the Nd-Fe-B/a-Fe
Nd-Fe-B/ex-Fe
multilayer films, the d Fe and dNd-Fe-B dependence of magnetization is roughly in
accordance with the estimation based upon the simple superposition of
magnetization of the Nd-Fe-B and ex-Fea-Fe phases.

--
I
obs. I __
I

E
~ 0
0
~

-11=--_
-I _- - -

--{).5 --{).25 o0 0.25 0.5


I-4JH(T)
JioH(T)
(a)
2
calc.
- I
E0 I
----1---
~
I
-I I
-2
-2':-::----::-'-::-.,,-------c--------::-'::-::--____,.'
--{).5 --{).25 0 0.25 0.5
/loH(T)
JioH(T)
(b)

Figure 7. 11 ( a) Observed and (b) calculated magnetization curves for glass /


Ti (30 nm)/ [a-Fe
[ex-Fe (5 nm) / Nd-Fe-B (30 nm)]5/ a-Fe
ex-Fe (5 nm) / Ti(30 nm) sample. It
should be noted that calculated minor loop overlaps with the curve of major loop where
demagnetization field is less than about O. 3 T.

In order to discuss the present results, a micromagnetic calculation of the


Nd-Fe-B/a-Fe multi layers has been performed by
magnetization curves for Nd-Fe-B/ex-Fe
using a Landau-lifshitz-Gilbert equation. First, the Nd-Fe-B/ex-Fe
Nd-Fe-B/a-Fe bilayer
lOxx 10
structure divided into 10 lOxx N,
N z x (N,
(N z is the number of cell along the z-
direction for micromagnetic calculation, N, =
N z = 1 - 12 depending on the layer
thickness) cells has been considered, as shown in Fig. 7. 14. The dimensions
of each cell were equal to those of a crystal grain for x, y directions (within
the film plane) and 5 nm for z direction (normal to the film plane). A periodic
boundary condition for x, y and z directions has been adapted to form
282 Hiroaki Kato et al.

obs.

-I

--{l.5
-0.5 --{l.25
-0.25 o 0.25 0.5
J.loH(T)
lloH (T)
(a)
2
calc.

-I

-2 ~====::'---f..-
-21:::=====::"'---..+--
--{l.5
-0.5 --{l.25
-0.25 o 0.25 0.5
lloH
J.loH(T)
(T)
(b)

Figure 7. 12 (a) Observed and (b) calculated magnetization curves for glass /
Ti (30 nm) / [ex-Fe (30 nm) / Nd-Fe-B (30 nm)]5/ ex-Fe (30 nm) / Ti (30 nm) sample.

multilayer without surface boundary. A total energy of the i-th


i -th cell for
micromagnetic calculation is assumed to be expressed by
6
= K;sin
E; = K ;sin2 S;
0; - ~
2:= 2dJ;kM; M k - 2M; H
2dJikMi (7.7)
k= 1 ;k
ik

where the first and third terms represent a magnetocrystalline


magnetocrystall ine anisotropy
energy and Zeeman energy, respectively, in which K K;i is the uniaxial
anisotropy contant for the i- th cell, and S; 0; is the angle between the easy axis
and magnetization direction in the i-th i- th cell. The easy direction of each Nd-Fe-
B grain is assumed to be randomly oriented, while the neighboring grains are
coupled with each other through interface with exchange interactions, defined
by the second term in Eq. (7. 7). The intergrain coulping constant J;k is
defined as an exchange energy per unit area, and d;k is a distance between
the center points of i-thi -th and k-th
k -th cells. Since the external field is applied
appl ied
parallel to the film plane, magnetostatic energy was certified to be negligible.
Three kinds of exchange interactions between the cells have been considered,
in which two intralayer coupling constants J hh (between cells of Nd-Fe-B) and
ex-Fe) are fixed to be 2.0 x
J ss (of a-Fe) 1O- 2 J/m2 , which is comparable to the value
X 10-
adopted by Fukunaga et al. (1996). The parameters of exchange constant J hs
High-Field Investigations on Exchange Coupling. . . 283

0.6 dNd-Fe-B=30 nm
obs.
o Jhs=O.IIXJ
Jhs=O.11 XJhh
o J hs=0.22XJhh
E:
E 0.4 l!. J hs=0.56XJhh
'"
:cr
~~
0.2

0 10 20 30
d Fe (nm)
(a)

dFe=IO nm
0.4

E:
E
~j 0.2

20 40 60 80 100
dNd-Fe-B (nm)
(b)

E:
E 0.2
~
j:::

o 10 20 30 40 50
d(nm)
(c)

Figure 7. 13 Dependence of the coercive field He on the layer thickness d in Nd-Fe-B/o:- Nd-Fe-B/a-
Fe multilayer films. (a)Ca) as a function of d F with fixed dNd-Fe-B = 30 nm, (b)
Cb) as a function of
= 10 nm, and (c)
dNd-Fe-B with fixed d F = Cc) as a function of d = = d F == dNd-Fe-B' Solid and open
circles represent the observed and calculated data, respectively, while solid lines are
guides to the eye.

(between Nd-Fe-B and ex-Fe cells) and crystal grain size d ery were adjusted so
as to best reproduce the experimental data. We found that calculated He
shows a maximum for dNd-Fe-B = dd ery (dNd-Fe-B is the thickness of Nd2 Fe14 B layer) .
284 Hiroaki Kato et al.

10

f!lo~~~~~mmEJJ
Nd-Fe-B
~ layer
Nd-Fe-B
layer
Ja-Fe
u-Fe layer

Magnetic field, H

Figure 7. 14 Schematic model of the present micromagnetic calculations


calculations. White and gray
boxes denote the hard (Nd-Fe-B) and soft (a-Fe) grains, respectively. External magnetic
field H is applied parallel to the layer (x - y plane).
plane) .

From this fact and the experimental results in Fig. 7. 13b,c, 13b, c, we have
determined the magnitude of d ery to be 15 nm. Figures 7. 11 band 7. 12b show
the calculated magnetization curves, corresponding to the films of [a-Fee d Fe ) /
nm ) J5 with a thickness of a-Fe layer d Fe = 5 and 30 nm,
Nd-Fe-B ( 30 nm)
respectively, in which the inter layer coupling constant was J hs = O. 11 X J hh . It
interlayer
should be noted that calculated minor loops show the spring-back behavior in
both small and large d Fe films, which are in agreement with the experiment.
Open symbols in Fig. 7.13 show the calculated He using the various magnitudes
of J hs parameter. Observed He dependence appears to be better reproduced
by the calculation when J hs = O. 11 X J hh . These results suggest that the
interlayer exchange-coupl ing strength (corresponding to J hshS)) is about an order
of magnitude smaller than that of the intralayer couplings (corresponding to J hh
or J SS).
55)' It should be noted, however, that not all the atoms at the interface of
hard and soft phases might be exchange coupled. The effective coupling
strength between soft and hard phases is the product of the coupl ing constant
per pair at the interface and the number of the coupled pair.

7.4.2 Nanodispersed Nd2Fe14B/a-Fe


NdzFe14B/a-Fe Films

As mentioned above, we have demonstrated the magnetic properties of


multilayered hard/soft nanocomposite magnets. In this system the coupling
between the two phases appears to be not sufficient, owing not only to the
smaller J hs
hS values but also to the smaller interface area arising from the

multilayer geometry. Nanodispersed geometry, on the other hand, has the


advantage of bringing in a larger interface area between hard and soft grains.
Here, we describe the results in such a system.
Thin films were deposited by RF-magnetron sputtering onto glass
High-Field Investigations on Exchange Coupling. . . 285

substrates in an Ar atmosphere. The base pressure of the sputtering system


was below 4 x 10 1O-- 55 Pa and the Ar pressure during the sputtering was 0.7 O. 7 Pa.
Thin films of Nd2 zFe14 B/ex-Fe system have the form of glass / Ti (100 nm) / Nd-
Fe-B (500 nm ) /Ti (100 nm )).. Since as-sputtered Nd-Fe-B phase was
confirmed to be in an amorphous state, it was annealed at T aa = 923 K for
10 min to crystallize the Nd2Fe14B
NdzFe14B phase. Volume fraction of the ex-Fe phase in
films, VFe' was adjusted by putting different numbers of ex-Fe chips (314 mm2 Z
))
with a fan shape on the Nd 13 Fe70 B 17 17 disk target with a diameter of 76 mm. The
composition of these films was determined by EPMA. The magnitude of V Fe ,
was estimated by taking both CD Rietveld analysis (Izumi, 1993) results of X-
ray diffraction patterns with Cu Kex radiations and (ll (Z) temperature dependence
data of remanent magnetization into consideration. Magnetization curves were
measured at room temperature by two kinds of vibrating sample magnetometer
systems, with the maximum applied fields of 16 and 140 kOe, respectively.
Ferromagnetic resonance experiments (Koyama et al. , 2000) were performed
by using the microwave frequency of 50 - 135 GHz in the fields of up to
150 kOe.
Figure 7. 15 shows the X-ray diffraction patterns for nanodispersed
Nd2Fe 14 B/ex-Fe films. Diffraction peaks of isotropic Nd
NdzFeI4B/ex-Fe 2Fe14 B and ex-Fe phases
NdzFe14B
are observed in addition to the underlayer and overlayer Ti and Ti-oxide
peaks. The relative intensity of ex-Fe reflection with respect to that of
Nd2 zFe14 B phase increases with increasing VFe' We then estimated the mean
grain size d ery of both phases by using the full-width at half maximum (FWHM)
data of diffraction peaks and Scherrer's formula, which are plotted against VFe
in Fig. 16. Inoue et al. (1995) reported that crystallization temperature Teryof
ex-Fe is lower than that of Nd2 zFe14 B in melt-spun nanocomposite ribbons. In the
crystallization process of Fe-rich samples, therefore, volume fraction of
nucleated ex-Fe crystallites is larger, which prevents the free growth of each
grain, resulting in the smaller ex-Fe grain size. Since the resultant average
distance between the ex-Fe grains is small in this case, subsequent nucleation
and growth of NdzFe14B
Nd2Fe14B crystallite will be limited.
Next we show in Fig. 7.17 7. 17 the magnetization curves of a series of
NdzFeI4B/ex-Fe films with different V~es. The external field was applied within
Nd2Fe14B/ex-Fe
the film plane. Although an existence of both soft and hard magnetic phases
was confirmed by the X-ray measurements, all the hysteresis loops show a
single-phase-like shape. Moreover, we observed a typical recoiling behavior
(Ishizone et aI., 1999) with high permeability in the minor loops. These
results thus indicate that Nd2 zFe14 B and ex-Fe grains are firmly exchange-
coupled. The coercivity He is quite high in the film with VFe = 5%
V Fe = 5 % and
decreases with increasing VFe' Reduced remanence M,/ M r / M s'
s , on the other
hand, becomes larger in high V VFe
Fe films.
In order to analyze these data, a micromagnetic calculation similar to that
mentioned above has been made (Ishizone et al. , 2000). A system of lOx
lOX 10 cubic cells is considered, in which each cell is assumed to represent a
286 Hiroaki Kato et al.

Nd 2 Fe l4
l4 B
o a-Fe
'"I> Ti
Ti

~
::i r--~v'-_-.J'J"'-'

~
o
'~
~
22%
.
c: r---'V\....,.....J''vV
o

42%


I>

68%

30 35 40 45 50
2(W)
21W)
Figure 7. 15 X-ray diffraction patterns for nanodispersed Nd,Fe14B/o:-Fe
Nd2Fel. Bfa-Fe films,
films.

60

E
s5
i!:l
1:l
.iii
';;;
40 0
~~

, 0

c:
.'~
OJ)

c:
20
"
(<l)
J)
Nd 2Fe l4B
~
o a-Fe

o0 20% 40% 60% 80% 100%


VFe
Vee
Figure 7.16 Mean size of Nd,Fel,B
Nd2Fel4B and a-Fe
o:-Fe grains estimated from the X-ray data,
data.

hard or soft magnetic grain,


grain. All the spins inside a cell are assumed to be
identical, namely homogeneously rotated.
rotated, The total energy of the system is
assumed to be the sum of anisotropy, exchange and Zeeman energy terms,
Eq, (7.
which are similar to those in Eq. 7). For computational economy, dipolar
(7,7),
interactions are ignored, since they are considered to give an unimportant
contribution (Fukunaga and Inoue, 1992), owing to the large anisotropy of the
HighField Investigations on Exchange Coupling. . .
High-Field 287

obs cal
VFe=5%
V
,---- -(
/. /~r
I--
100-

o r -
-1
j l.--/ .-1
~J _ V
J=2
1 10%
(T
('

o IT
-1I
7 IJ u J 11
1 22%
( 7 ( T
T

-1
7 [7 lJ J-9
1 42%

o // f7 f
-11
~JJ
--lIJ U J-5
1I 68% ~/
~/

o
)) J=5
-1
-2 -1
-I 0 -1 0 2
~H(T)
}JoH(T) ~H(T)
}JoH(T)

Figure 7. 17 Observed and calculated magnetization curves for nanodispersed Nd2 Fe1.
Fe14 S/
ex-Fe films. Magnetization M is normalized by saturation magnetization M s ,' which was
defined as M at 10 T.

system. The grain size is assumed to linearly depend on V Fe by using the


experimental data given in Fig.7.16.
Fig. 7.16. For simplicity, only the results for the
case of fixed exchange coupling constant (J = = J hh == J ss == J hshS ) are presented
here.
Magnetization curves were thus calculated with different values of the
exchange-coupling constant J. We have found that the shape of the loops
qualitatively changes at the critical exchange-coupling value of J = = J c .' That
is, for J < J c ' loops are two-phase, showing two distinct decreases of
magnetization owing to the independent switching of soft and hard grains,
while for J ~ J c the double step demagnetization process is merged into a
single one owing to the strong coupling. Calculated curves where J = J c are
given in Fig. 7.17 in comparison with the experiments. A general trend of the
observed hysteresis loops is well reproduced by these calculations. The values
of J c thus adopted are plotted against VFe in Fig. 7.18. The magnitude of V Fe is
very small in the sample with VFe = 5 %, exhibits a maximum around VFe " -
10 %, then gradually decreases with increasing V Fe' Since J is defined in
terms of energy per unit interface area, it should be constant irrespective of
V Fe if the interface or intergrain states remain unchanged.
288 Hiroaki Kato et al.

15
M vs.
YS. H
o FMR

o 20% 40% 60% 80%


VFe

Figure 7. 18 Exchange-coupling constant J as a function of ex-Fe volume fraction VFe for


VFe

nanodispersed Nd2Fe'4
Nd2Fe14 B/ex-Fe films. Closed circles denote the values determined from
micromagnetic analysis of magnetization curves, while open squares denote the values,
determined from FMR experiments.

Therefore, it is necessary to confirm this V Fe dependence of J by using


independent experimental technique such as ferromagnetic resonance (FMR).
In a study of magnetic multilayers,
multi layers , FMR has been a powerful technique to
retrieve information on the interlayer exchange-coupling (Heinrich and
Cochran, 1993). As for a hard/soft nanocomposite system, it is also possible
to apply this technique to estimate the intergrain exchange coupling.
However, it is not easy to observe an FMR signal in a hard magnet such as
Ndz Fe'4
Fe14 S,
B, since the resonance field and frequency become very large owing to
its strong magnetic anisotropy. We made an FMR measurement by using a
high-frequency (f = = 50 - 195 GHz) and high-field (up to 100 kOe) electron
spin resonance (ESR) apparatus to investigate the exchange coupling in a
Ndz Fe'4
Fe14 S/a-Fe
B/ex-Fe system (Kato et al., a!., 200
20011b).
b). Film samples were set in a
cylindrical resonant cavity so that the external dc magnetic field was
perpendicular to the film plane. We observed two major absorption peaks
except for pure a-Fe
ex-Fe films ((V = 100 % ), as shown in Fig. 7. 19. As can be
V Fe =
seen from this figure, the position of lower-field peaks are almost independent
of V Fe' while the higher-field ones shift toward higher resonance position with
increasing V Fe' We then made a model calculation of resonance field versus
frequency relation in order to analyze these data. For simpl icity, a two-
sublattice model is adopted, in which each sublattice represents a hard or soft
phase. Assuming that the easy magnetization axes are within the film plane
(xz-plane), which is normal to the external field direction (y-axis), then the
total energy of the system can be written by

:8
E = ~ Vie
V i [ -- M ij Hsin8
i= h.s
jsini
Hsin8;sincf>i 21TMfsinz2 8
+ 2rrMfsin jsin z2 j
8;sin cf>i + K 2z
i sin 8
K;sin z2 cf>J
i sin ;]
8;sin
i=h.s
(7.8)

where 8
8;i and i
cf>i are defined as the angles between the magnetization vector
High.Field Investigations on Exchange Coupling. . . 289

135 GHz

22%

o 4 6 8 10
J1l1oH
oH(T)
(T)

Figure 7. 19 Fe,. S/
FMR spectra for nanodispersed Ndz2 Fe" B/ a-Fe
ex-Fe films taken at a fixed
frequency of 135 GHz.

M i and the z-axis, and x-axis, respectively. The subscript i denotes a hard or
= h, s. The first term multiplied by the volume fraction Vi
soft phase with i =
denotes the summation of Zeeman, demagnetizing field, and anisotropy energy
the i phase, whi
terms of the; while
Ie the second term represents the exchange-coupIing
exchange-coupling
energy between hard and soft phases, in which S is the interface area
between the two phases. In order to validate the two-sublaUice
two-sub lattice mapping of the
real system, we assumed that S is proportional to d~ry x V Fe ( 100 - V Fe)' by
adopting observed values of ddcry ery and V Fe' Moreover, external fields were
assumed to be sufficiently strong so that eh= = es ==90 and <Ph e
= <Ps
4>h = =90. Final
4>s =
relation between resonance frequency and fields was obtained according to the
procedure of Schmool and Barandiaran (1998). Figure 7.20 shows an example

200
obs. vFe
22%
o 68%
150

~
N
J:
2- 100
'-,

calc. VFe
-_.- 22% (J=2)
-- -- (1=2)
- - 22% (J=7)
- - 68% (J=2)
(1=2)
....68% (J=4)
68%

o 4 6 8 10
J1 oH(T)
l1oH (T)

VFe = 22 %
Figure 7. 20 Calculated and observed resonance frequency as a function of field for V
and 68% films. Values of exchange-coupling constant J are expressed in units of mJ/rrth.
290 Hiroaki Kato et al.

of observed and calculated resonance frequency plotted as a function of fields


for VFe = 22 % and 68 % samples. Although the degree of fitting is not
satisfactory, we can at least confirm that J for the VFe = 22 % sample is
significantly larger than that for the 68 % sample. The values of J thus
estimated are plotted in Fig. 7. 18, which are in agreement with those
determined from micromagnetic analysis of magnetization data.

7.5
7. 5 Conclusions

From the magnetization measurements and micromagnetic analysis for the


multilayer nanocomposite system, the interlayer exchange-coupling strength
J hshS is found to be about an order of magnitude smaller than that of the intralayer
coupl ings. These values are independent of the layer thickness, hence the
volume fraction of soft magnetic phase VFe' In the case of nanodispersed
films, both magnetization and FMR experiments showed that J does depend on
VFe and, for VFe> 10%, decreases with increasing VFe' The magnitude of J
is thought to be strongly affected by the interface state, since the coupling
originates from the exchange interaction between a pair of atoms at the
interface. If there exists an intergranular phase at the interface, J might
decrease considerably owing to its short range character. For example, an
existence of a Nd-rich amorphous phase was reported by Mishra (1986) in the
exchange-decoupled melt-spun Nd-Fe-B ribbons. In the case of exchange-
coupled magnets, there is no evidence of such intergranular phase and direct
neighboring of the grains were observed (Mishra and Panchanathan, 1994;
Zern et al., aI., 1998). Observed VFe-dependent variation of J in the present
nanodispersed films may thus be related to the thickness of an intergranular
phase at the interface of grains. Wecker and Schultz (1987) reported that the
crystallization temperature T cry cry of amorphous Nd-Fe-B ribbons increases with
decreasing Nd content. Since the annealing temperature T a was kept constant
in the present series of nanodispersed films, there is a possibility
possibil ity that volume
fraction of a Nd-rich amorphous phase increases with increasing VFe' and
hence with decreasing Nd content, owing to the reduction of temperature
difference T TaT
a T cry' Although not confirmed directly, an observed decrease in J
for the V VFe>
Fe > 10% region can be explained by such mechanism.
It should be noted that the magnitude of J hs hS for multilayerd magnets are

comparable to that in nanodispersed films with VFe '" -- 70 %. This result


suggests that, in the multilayer film, there exists a certain amount of interlayer
third phase between the hard and soft layers, which naturally reduces the
exchange coupling force. In the case of SmCo/Fe epitaxial bilayers, observed
magnetization curves were well reproduced by the J hs value which has
comparable magnitude to J hh and J ss (Fullerton et al., aI., 1998b). Since their
High-Field
HighField Investigations on Exchange Coupling...
Coupling. . . 291

values of J hh and J ss are comparable to the atomic exchange within each layer,
this result indicates that the interfacial coupling between different layers or
grains can be as strong as the interatomic exchange interaction unless the
intervening phase exists.
As compared with the intergrain exchange coupling in nanocomposite
magnets, we can obtain more systematic information on the magnitude of
exchange coupling J RFe between R-Fe or molecular field in R-Fe intermetallics,
as mentioned in Section 7. 2 and 7. 3, and also as reported in literatures
(Belorizky et ai., 1987; Radwanski et ai., 1990a). Variation of J RFe across
the lanthanide series has been understood at least qualitatively (Belorizky
et ai.,
al. , 1987). It seems that there still remain many problems to solve, as for
the intergrain exchange coupling. Knowing this mechanism will bring us to
manipulating or controlling the magnitude of the coupling so as to optimize the
magnetic properties of nanocomposite magnets.

References
Algarabel, P. A.A.,, A. Del Moral, M. R. Ibara and C. Marquina. J. Magn.
Magn. Mater. 114: 161 (1992)
AI-Omari, I.A. and D.J. Sellmyer. Phys. Rev. B 52: 3441 (1995)
Asti, G. and F. Bolzoni. J. Magn. Magn. Mater. 20: 29 (1980)
Belorizky, E., M. A. Fremy, J. P. Gavigan, D. Givord and H. S. Li. J.
Appl. Phys. 61: 3791 (1987)
Cadogan, J. M. , J. P. Gavigan, D. Givord and H. S. Li. J. Phys. F 18: 779
(1988).
Clark, A. E. Handbook on the Physics and Chemistry of Rare Earths. edited
by K. A. Gschneider, Jr. and L. Eyring. North-Holland, Chap. 15, p. 231
(1979)
Coehoorn, R. and J. Duchateau. Mater. Sci. Eng. 99: 131, (1988)
Coehoorn, R. , D. B. de Mooij. J. P. W. Duchateau and K. H. J. Buschow. J.
Phys. (Paris) C 8: 669 (1988)
Coehoorn, R. , D. B. de Mooij and C. de Waard. J. Magn. Magn. Mater.
80: 101 (1989)
Coey, J.M.D.
J. M. D. and Hong Sun. J. Magn. Magn. Mater. 87: L251 (1990)
Fukunaga, H. and H. Inoue. Jpn. J. Appl. Phys. 31: 1347 (1992)
Fukunaga, H., N. Kitajima and Y. Kanai. Mater. Trans. JIM 37: 864 (1996)
Fullerton, E. E.,
E. , J. Samuel Jiang, C. H. Sowers, J. E. Pearson and S. D.
Bader. Appl. Phys. Lett. 72: 380 (1998a)
Fullerton, E. E. , J. S. Jiang, M. Grimsditch, C. H. Sowers and S. D. Bader.
Phys. Rev. B 58: 12,193 (1998b)
Givord, D.,
D. , H. S. Li, J. M. Cadogan, J. M. D. Coey, J. P. Gavigan, O.
Yamada, H. Maruyama, M. Sagawa and S. Hirosawa.
Hi rosawa. J. Appl. Phys. 63:
3713 (1988)
Heinrich, B. and J. F. Cochran. Adv. Phys. 42: 523 (1993)
292 Hiroaki Kato et al.

Herbst, J. F. Rev. Mod. Phys. 63: 819 (1991)


Hirosawa, S. and M. Sagawa. Solid State Commun. 54: 335 (1985)
Hirosawa, S. , Y. Matsuura, H. Yamamoto, S. Fujimura,
Fuj imura, M. Sagawa and H.
Yamauchi.J. Appl. Phys. 59: 873 (1986)
Hirosawa, S. , H. Kanekiyo and M. Uehara. J. Appl. Phys. 73: 6488 (1993)
Inoue, A.A.,, A. Takeuchi, A. Makino and T. Masumoto. IEEE Trans. Magn.
MAG-31: 3626 (1995)
Ishio, S. and F. Sato. J. Magn. Soc. Jpn. 12: 259 (1988)
Ishizone, M. , A. Sakuma, H. Kato and T. Miyazaki. J. Magn. Magn Soc. Jpn.
23: 1105 (1999)
Ishizone, M. , T. Nomura, H. Kato and T. Miyazaki. J. Magn. Soc. Jpn.
24: 423 (2000)
Izumi, F. The Rietveld Method. ed. by R. A. Young. Oxford Univ. Press,
Oxford, Chap. 13 (1993)
Kato, H., M. Yamada, G. Kido, Y. Nakagawa, S. Hirosawa and M.
Sagawa. J. de Physique, 49: C8575 (1988)
Kato, H., M. Yamada, G. Kido, Y. Nakagawa, T. Iriyama and K.
Kobayashi. J. Appl. Phys. 73: 6931 (1993)
Kato, H. , D. W. Lim, M. Yamada, Y. Nakagawa, H. Aruga Katori and T.
Goto. Physica B 211: 105 (1995)
Kato, H., T. Ishizaki, F. Sato and T. Miyazaki. J. Magn. Soc. Jpn. 23:
495 (1999)
Kato, H., T. Nomura, M. Ishizone, H. Kubota T. Miyazaki and M.
Motokawa. J. Appl. Phys. 87: 6125 (2000a)
Kato, H., M. Ishizone, T. Miyazaki, K. Koyama, H. Nojiri and M.
Motokawa. Proc. 16 th Int. Workshop on Rare-Earth Magnets and Their
Applications, edited by H. Kaneko, M. Homma and M. Okada. The Japan
Insititute of Metals, p. 547 (2000b)
Metals,p.
Kato, H. , T. Ishizaki and T. Miyazaki. IEEE Trans. Magn. 37: 2702 (2001a)
Kato, H., M. Ishizone, T. Miyazaki, K. Koyama, H. Nojiri and M.
Motokawa. IEEE Trans. Magn. 37: 2567 (2001b)
Kneller, E. F. and R. Hawig. IEEE Trans. Magn. 27: 3588 (1991)
Koyama, K. K.,, M. Yoshida, H. Nojiri, T. Sakon, D. Li, T. Suzuki and M.
Motokawa.
Motokawa.J. J. Phys. Soc. Jpn. 69: 215 (2000)
Lim, D. W. , H. Kato, M. Yamada, G. Kido and Y. Nakagawa. Phys. Rev.
B 44: 10,0
10 ,0 14 (1991)
Lim, D. W. , H. Kato, M. Yamada, G. Kido and Y. Nakagawa. J. Magn.
Magn. Mater. 104-107: 1429(1992)
1429 (1992)
Mishra, R. K. J. Magn. Magn. Mater. 54-57: 450 (1986)
Mishra, R. K. and V. Panchanathan. J. Appl. Phys. 75: 6652 (1994)
Nakagawa, Y., Y. , H. Kato, D. W. Lim, G. Kido and M. Yamada. Proc. 6 th
Inter. Symposium on Magnetic Anisotropy and Coercivity in Rare Earth-
Transition Metal Alloys, ed. by S. G. Sankar. Carnegie Mellon University,
Pittsburgh, p. 12 (1990)
High-Field Investigations on Exchange Coupling. . . 293

Obermyer,
Obermyer. R. T. and F. Pourarian. J. Appl. Phys. 69: 5559 (1991)
Parhofer,
Parhofer. S. M.,
M . J. Wecker,
Wecker. C. Kuhrt,
Kuhrt. G. Gieres and L. Schultz. IEEE
Trans. Magn. 32: 4437 (1996)
Radwanski. R.
Radwanski, R.JJ.., J. J. M. Franse and R. Verhoef. J. Magn. Magn. Mater.
J.J.M.
83: 127 (1990a)
Radwanski. R. J.
Radwanski, J ., R. Verhoef and J. J. M. Franse. J. Magn. Magn. Mater.
83: 141 (1990b)
Sagawa,
Sagawa. M.M ., S. Fujimura,
Fujimura. N. Togawa,
Togawa. H. Yamamoto and Y. Matsuura. J.
Appl. Phys. 55: 2083 (1984)
Schmool, D. S. and J. M. Barandiaran. J. Phys.: Condense. Matter 10: 10,
Schmool. 10.
679 (1998)
Shindo,
Shindo. M. Ishizone. H. Kato.
M ., M. Ishizone, Kato, T. Miyazaki and A. Sakuma. J. Magn.
Magn. Mater. 161: Ll L1 (1996)
Shindo,
Shindo. M.
M ., M. Ishizone,
Ishizone. A. Sakuma,
Sakuma. H. Kato and T. Miyazaki. J. Appl.
Phys. 81: 4444 (1997)
Sidorov,
Sidorov. V. A. and L. G. Khvostantsev. J. Magn. Magn. Mater. 129: 356
(1994)
Skomski, R.,
Skomski. R . J.M.D.Coey. Phys. Rev. B 48: 15,812
B48: 15.812 (1993)
Wecker, J.
Wecker. J., L. Schultz. J. Appl. Phys. 62: 990 (1987)
Withwanawasam. L.,
Withwanawasam, L . A.S. Murphy,
Murphy. G.C. HadjipanayisandR.F. Krause. J.
Appl. Phys. 76: 7065 (1994)
Yamada,
Yamada. M.M ., H. Kato,
Kato. H. Yamamoto and Y. Nakagawa. Phys. Rev. B 38: 38 :
620 (1988)
Zern. A.,
Zern, A . M. Seeger,
Seeger. J. Bauer and H. Kronmuller. J. Magn. Magn. Mater.
184: 89 (1998)

The authors would like to thank Dr. A. Sakuma, M. Shino, D. W. Lim, Ishizaki, T.
Lim. T. Ishizaki.
Nomura and M. Ishizone for their collaboration. We are grateful to Professor M. Yamada for
his fruitful discussion about the crystal-field calculations. Thanks are also due to Professor Y.
Nakagawa, Professor G. Kido and staff members of High Field Laboratory for
Superconducting Materials. This work was partly supported by the Murata Science
Foundation, by a Grant-in-Aid for Scientific Research (No. 09650002) from the Ministry of
Foundation.
Education, Science, Sports and Culture, by the Mazda Foundation and by Iketani Science and
Technology Foundation.
8 Fabrication and Magnetic Properties of
Nanometer-Scale Particle Arrays

S. Wirth and S. von Monlnar

8. 1 Introduction

Nanometer-scale magnetic particles attract much interest since they can be


used for testing basic concepts of ferromagnetism and, possibly, for
appl ications in high density magnetic storage. These goals, however, call for
applications
extremely small particles which are well shaped and arranged. Nanometer size
magnetic particles require both suitable manufacturing techniques as well as
sophisticated measuring techniques capable of detecting very small magnetic
moments. Provided that such techniques are available, the magnetization
behavior of small individual particles may be studied Here, a detailed
understanding of the magnetization reversal mechanism is of special merit.
After a brief introduction of the fabrication and measuring techniques for
small magnetic particles, we will present the methods used by us: scanning
tunneling microscope assisted chemical vapor deposition and Hall
gradiometry, respectively (in addition, magnetic force microscopy will be
discussed to a limited extend). Measurements on arrays of well shaped and
aligned nanoparticles allow us to study the magnetization reversal in these
particles, the effect of thermal activation, and interaction effects. Here, we
concentrate on the classical regime for magnetization reversal in a temperature
range from 10 - 300 K. A phenomenological nucleation-type model for the
magnetization reversal in small elongated particles is supported even though
the particles of the different arrays are only 9 - 20 nm in diameter. Some of
the results are compared to numerical simulations. Our results indicate that the
uniform rotation model is applicable for reversible rotation but not for the
magnetization reversal. The latter is important for understanding these
complex mechanisms as well as for applications.
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 295

8. 2 Fabrication of Regularly Arranged and Shaped


Particles on a Nanometer Scale

8. 2. 1 Overview

A number of new fabrication procedures has been developed to produce arrays


of ever smaller magnetic particles. Here, the magnetic particle or grain size
should be well below the single domain size of the specific material. In
keeping with traditional procedures, Iithographic
lithographic methods have been
improved, not only to make a small number of magnetic nanoparticles but also
to fabricate single-domain particles of regular shape and arrangement over
areas of the order of 1 cm 2 . Features much smaller than with photolithography
were obtained by using electron beam lithography (Smyth et al. , 1991; Chou
et al., 1994a, 1994b; New et ai., aI., 1994, 1995; Ruhrig
ROhrig et ai.,
aI., 1996a;
Cowburn et al.,
aI., 1997; Martin et al.,aI., 1998; Wong et al.,aI., 1999; Dunin-
Borkowski et aI.,
al. , 1999; Haginoya et aI.,
al. , 1999). The actual patterning of the
magnetic material is often performed by reactive ion etching or ion irradiation
(Chappert et al.,
aI., 1998). The writing of structures into the resist of each
sample, however, is quite timeconsuming, especially if large areas are to be
written. Maskless patterning by steering focused ion beams (Nakayama et
al., 1998) suffers from the same deficiency. Therefore, patterns were
produced by stencil masks (Terris et aI.,al., 1999) that can be re-used several
times. The ions can either be used to remove unprotected areas of the
magnetic material or to locally change its magnetic properties (Saifullah
et al. , 1995; Terris et al. , 1999).
Nanometer-scale particle arrays were also fabricated by X-rays
(Rousseaux et al. , 1995; Smith et al. , 2000). Moreover, feature sizes down
to about 100 nm could be obtained by interference (holographic) lithography
(Fernandez et aI., 1996; Kirsch et aI., al., 1997; Kreuzer et aI., 1998; Savas
et aI., 1999; Haast et aI., 1999). Here, a laser beam is split and its
interference pattern is used to expose a resist. With only two perpendicular
exposures, square arrays of particles can be defined over several square
centimeters. The same principle can be applied for direct writing by using the
heat input of the interference pattern to locally form Co nanodots out of a Co-C
precursor material (Zheng et al. , 2001).
Rather than using an energetic beam to locally modify the resist, it can be
patterned by using a sharp tip. With this nanoimprint lithography, structures
down to 25 nm were fabricated (Chou et al., al. , 1996).
The nanostructures to be fabricated are less likely to be damaged if the
296 S. Wirth and S. von Monlnar

deposition process follows the patterning. Again, resist masks as well as


stencil masks for repetitive use (Park et ai.,
aI., 1997) have been employed.
Here, electrodeposition has become an attractive method for filling nanoporous
materials since systems with changing compositions can be deposited (Meier
et al. , 1996; Zangari and Lambeth, 1997; Schwarzacher et al. , 1997; Duvail
et al. , 1998).
A less time consuming approach is to define the particles directly while
depositing the magnetic material. Thin films consisting of nanometer-size
grains have been fabricated from materials of high coercivity, e. g. SmCo
(Lambeth et ai.,al., 1996), FePt (Li and Lairson, 1999) and .CoPtCoPt (Ichihara
et al., 1998; Stavroyiannis et ai.,
aI., 1999; Yu et ai.,
aI., 1999). In continuous
films, however, the grains are magnetically coupled and one bit is written onto
more than one grain. This deteriorates the signal-to-noise ratio. Separated
individual particles can be produced by cluster beam deposition (Billas et al. ,
1993) or island growth of very thin films (Chambliss et al. , 1991). Moreover,
structured substrates, e. g. , step edges (Patel and Pepper, 2000) or colloidal
templates (Li et al. , 2000) can be used. A very promising approach is the
synthesis of monodisperse FePt particles with a tunable diameter from 3 -
10 nm (Sun et al., 2000). These nanoparticles self-assemble and, upon
anneal ing, transform into ferromagnetic nanocrystal assembl
annealing, ies that support
assemblies
high-density magnetization reversal transitions.

8.2.2 Scanning Tunneling Microscope Assisted Chemical Vapor


Deposition
Scanning tunneling microscopy (STM) is capable of atomic resolution (Binnig
et al., 1982). Therefore, research effort was put into the STM-based
fabrication of features and surface modification on a similar length scale. The
possibility to deposit metallic particles from a variety of vaporous
organometallic precursors was recognized early (Ehrichs et ai., aI., 1988;
McCord and Awschalom, 1990).
The iron particle arrays under investigation here are fabricated by a
combination of chemical vapor deposition (CVD) and STM (Kent et al al..,
1993; Wirth et al.
al.,, 1998a). In principle, a commercially available STM with
three modifications is used. First, a comparatively high tip bias voltage is
appl ied (usually - 17 V, see below). Second, a vaporous organometall ic
compound can be introduced into the vacuum system at a moderate pressure.
Due to the tip bias, an electrical field is created within which the iron
pentacarbonyl decomposes and a deposit is grown at the tip position. As the
deposit grows, the tip is retracted by keeping the tunneling current (and
hence, the separation between tip and the deposit's top) constant. The actual
deposit height can therefore be monitored via the STM feedback loop. When
the particle has grown to the desired height, the tip is retracted completely
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 297

and moved to the next location on the sample surface where the process is
repeated to form the particle array. Hence, the position of the particles with
respect to each other and to any feature on the sample surface can simply be
controlled by steering the tip. The growth process and the tip steering are-
as a third modification - computer controlled. All STM tips are cleaned in-situ
by resistive heating and a field-emission process before usage.
Since the particles are grown one by one the fabrication process is too
time consuming (about 10 s per particle) for magnetic storage applications.
Processes in which a large number of particles is grown at once by using
multiple tips could be envisioned but have not yet been realized. The particle
arrays as grown so far, however, are perfectly suitable to address physical
problems especially since the particles can be positioned precisely and
individually, and their height can easily be controlled. At present, particles
with heights between 50 and 250 nm have been grown (Wirth et al., 1998a,
1999).
As a morphological restriction, the interparticle spacing must not be
smaller than the particle height due to the V-shaped tip. Otherwise clustering
of the particle that might be used to write continuous lines or other extended
feature occurs (Rubel et al. , 1994).
The mean particle diameter is mainly determined by the tip bias voltage
and the precursor pressure. The former also influences the quality of the grown
particles and is therefore kept at - 17 V. At this voltage, only occasionally
are minimal deposits found around the particles. The higher the precursor
pressure the faster and narrower the pillars grow. Arrays of optimal
homogeneity have been grown for pressures ranging from 20 - 26 IJ1..1 Torr
resulting in particle diameters d between 9 and 20 nm. For lower pressure the
growth rate is too slow and some deposit is found surrounding the particles,
probably due to diffusion of the adatoms on the surface (Kent et al. , 1993). In
addition, we cannot exclude the influence of the specific tip on the grown
particle diameter. We found some variance of the particle diameters for
identical growth parameters if different tips were used. It should be noted that
all quoted diameters are those of the magnetic core of the particles. These
cores consist of bcc iron (as revealed by TEM) (Kent et aI., 1993) and are
surrounded by a carbon coating. Even though this coating decreases the
magnetic volume of the particles, it also reduces the oxidation and aging of the
samples under air. Our oldest sample did not show any deterioration of its
magnetic properties after three years in air.
An analysis of the magnetic properties of particles grown under optimized
conditions indicates that the particles typically consist of a single or very few
grains. From this, we believe our grain sizes to be considerably larger than
those revealed by early TEM investigations (-9 ( ....... 9 nm) (Kent et al.,
aI., 1993).
The following restrictions apply to substrate surfaces: CD They must be
~ The surface must be sufficiently smooth.
conductive to perform tunnel ing; CZ>
These restrictions leave quite a variety of growth capabilities as seen in Fig.
298 S. Wirth and S. von Monlnar

8. 1. So far,
far. we have grown onto gold (non-magnetic),
(non-magnetic). permalloy (soft-
magnetic) and,
and. to a limited extent.
extent, onto niobium (superconducting at low
Furthermore. Kent et al. (1993) has grown directly
temperatures) surfaces. Furthermore,
onto Si. Different diameters and particle arrangements have been realized.
The square shape of the bigger particle's cross-section is Iikely
likely to be caused
by the bcc iron structure (Fig. 8. 1c,e).
The precursor gases used in the general CVD process can be decomposed
by heat or by energetic electrons.
electrons, ions or photons. Hence, one might presume
that the field-emitted electrons cause the decomposition of the precursor (de
Lozanne,
Lozanne. 1994). Indeed,
Indeed. the growth characteristics of the particles seem to
point toward electron-induced decomposition (Kent et al., 1993). A field-
induced decomposition component may also be present. This. This, however.
however, may
result in non-magnetic deposits of high carbon content.

l/lm
I 11m

(a)
.. .
-~~~
.. ~ ..
_., ...
- ~ .. ~
p

.
., '
, . . .. . .. . ..
"'.' " '.'
.. . . . . . ., ... . ...................
~
.. OJ
oj ..........,...

- '.
. . . . . . .... .. ...
~ .
. .,
, . . .
..
,... ....... '" ,
.. .. ..... .. .. ..... ... .. ..... .. .. .
'
.. .. + " .. .. .. ..
. .. . . .. ..... . . .....
~
,<; ~

.. . . ..
(b)
(d)
y
T

,
I
a

.
(c)

........


(e)
~ .
Figure 8.1 Arrays of particles grown by 8TM assisted CVD. (a) - (c) growth onto Au;
(d),
(d) , (e) growth onto permalloy. Different particle diameters and arrangements can be
realized.

Changing the sign of the tip bias voltage causes the deposit to grow onto
the tip (Kent et aI.,
al . 1993). This could open up the possibility to fabricate
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 299

magnetic force microscopy (CMFM)


MFM) tips with improved spatial resolution by
putting a particle on top of a regular atomic force microscopy(AFM)
microscopyCAFM) tip. The
versatility of our fabrication method may be enhanced even further by using
CRubel et al. , 1994) and
different metal carbonyls. Here, nickel tetracarbonyl (Rubel
dicobalt octacarbonyl are the precursors of choice if magnetic applications are
of interest. As an additional advantage of our fabrication technique, the
particles can be grown at a predefined position on the substrate. Towards this
end, the STM is used in regular image mode. If the feature of interest is found
on the substrate, the growth is initiated at that exact position.

8. 3 Magnetic Measurements on Extremely Small Particles

8. 3. 1
8.3.1 Overview
Closely intertwined with the development of fabrication methods is the
improvement of measuring capabilities for even smaller particles. Especially in
cases where only a small number of particles is fabricated or is to be
investigated, the measuring techniques need to be able to cope with even
smaller magnetic moments. In these cases, methods that provide locally
resolved magnetic information on the length scale of the particles or better are
preferable. Scanning probe microscopes (SPM) CSPM) use a microscopic sensor or a
focused beam and measure a synchronous signal from sensor and/or sample
that is magnetic in origin (forCfor reviews, see, e. g. (Dan CDan Dahlberg and
Proksch, 1999; de Lozanne, 1999.
One of the most popular magnetic imaging techniques is near field MFM
due to its high spatial resolution, versatility and natural combination with AFM.
A vibrating magnetic tip is scanned over the sample surface and changes in its
vibrating motion due to magnetic interactions between tip and sample are
recorded. Spatial resolution down to 20 nm has been reported (RugarCRugar et al. ,
CGrUtter et al. , 1995; Porthun et al. ,
1990). MFM was reviewed extensively (Gri.itter
1998; Hubert and Schafer 1998; Proksch, 1999) and will therefore be
discussed only briefly and with respect to our specific application in Section
8.3.4.
One of the disadvantages of MFM is the potential invasive action of the tip
due to its magnetic stray field (Thiaville
CThiaville et aI., 1997). In case of Hall
magnetometry, a non-magnetic sensor detects magnetic fields by measuring
Hall voltages. Microfabricated sensors made of a 20 electron system (2DES) C2DES)
allow local measurements on magnetic (Lottis CLottis et aI., 1992; Kent et al. ,
1994; Ye et al. , 1995) and superconducting materials (GeimCGeim et aI., 1997a).
It is also used in the present study (see
Csee following sections). Scanning Hall
300 S. Wirth and S. von Monlnar

probe microscopy (Hallen et al., 1993; Oral et al., 1996; Schweinbbck et


al. , 2000) can reach resolutions as low as O. 25 ~m with superior sensitivity
compared to MFM. The temperature and field range of use as well as the
sensitivity depend mainly on the 2DES material (for a review, see (Bending,
1999. Note that quantitative investigations are simpl ified if the Hall sensor is
used in the ballistic regime (Peeters and De Boeck, 1999).
The most sensitive sensors for magnetic flux-capable of measuring single
flux quanta-are superconducting quantum interference devices (SQUIDs).
Micro-SQUIDs have been fabricated by patterning niobium films to form
Josephson junctions (Wernsdorfer, 1995a). With these, excellent experiments
on the magnetic behavior of small single particles have been performed
(Wernsdorfer, et al.,aI., 1997). SQUIDs are non-invasive but suffer severely
from temperature and field limitations to maintain their superconducting state
and from considerable experimental complexity. Scanning SQUID
magnetometry with a spatial resolution of about 10 ~m was used to study the
flux distribution in superconductors (K irtley et al., 1995, 1997) in
unprecedented detail.
Another non-invasive technique is Kerr microscopy where polarized light
is used to probe the sample magnetization (for a review, see (Hubert and
Schafer, 1998. A rotation of the polarization can be detected in reflection or
transmission (Faraday effect). The magnetic and optical properties of a
ferromagnet are connected via the spin-orbit coupling that couples electron
spin and its motion. Hence, the effect is best observed in rare earths, but
digital image difference techniques allow investigation of almost all magnetic
materials (Qiu and Bader, 1999). Spatial resolution is Iimited to the order of
100 nm. It can be conducted in applied fields and can be combined with MFM
(Zueco et aI., 1998). The strength of Kerr microscopy is certainly its high
temporal resolution well below 1 ns. Scanning Kerr microscopy allowed the
spatio-temporal investigation of magnetization switching (Hiebert et al.,
1997; Ju et al. , 1999; Choi et al. , 2001).
The secondary electrons generated in a scanning electron microscope
(SEM) keep the spin polarization information of their initial state. Hence, a
polarization analysis (scanning electron microscope with polarization analysis,
SEMPA) of these electrons can reveal information on the sample magnetization
at its surface. High spatial resolution (about 10 nm) and quantitative analysis
rely on the phase modulation of transmitted electrons (Lorentz microscopy)
passing through a region of magnetic induction, i. e. a magnetic sample (see
(Chapman and Scheinfein, 1999) for a review). Both structural (with
resolution below 1 nm) and magnetic information can be obtained. The
objective lens must be modified to circumvent its magnetic field but can be used
for applying a field to the specimen. Magnetization reversal in small elements
has been investigated (Mankos et aI.,
al. , 1994; Thien Binh et al.,
al. , 1998; Kirk
et al. , 1999). Quantitative analysis can be performed by electron holography
( Hirayama et al. , 1993). A drawback is extensive sample preparation.
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 301

X-ray magnetic circular dichroism is a very sensitive tool: element


specific spin and orbital magnetic moment can be measured. In magnetic
matter, the absorption of the x-rays depends on their polarization. However,
considerable experimental setup is involved in using synchrotron radiation.
Investigations of small particles are sparse (Ohresser et al. , 2000).

8.3.2 Hall Gradiometry

Quantitative results for the total magnetization of our particle arrays are
obtained by Hall gradiometry. These measurements rely on the particles'
magnetic stray field through a 2DES. This stray field is about 1 mT or less,
much smaller than the particles' switching or anisotropy fields. It is
necessary, therefore, to compensate the Hall voltage induced by the external
field in the Hall cross carrying the array by an empty, adjacent cross on the
same substrate. This results in a cancellation leaving only the excess signal
due to the array's magnetization (Kent et al., ai., 1994; Wirth et ai.,
aI., 1998a).
Such a Hall gradiometer configuration is depicted in Fig. 8.2. In principle, the
currents through the two Hall crosses, /1 and /2' should be equal in magnitude
but opposite in direction. However, slight inhomogeneities in the fabrication of
the crosses may lead to minute differences in their Hall responses which are
nulled out by adjusting these currents. Our ill - V semiconductor Hall crosses
are prepared by photolithography and wet chemical etching of the substrate
(GaAs-Ga
(GaAs-Ga07 Al o3
07 AI 03 As, typical values for carrier concentration and mobility are
n2D=3.2X 10 11 cm- 2 and /J(30
n2D=3.2x /..1(30 K)=4.5x 10 5 cm 2 /(V' s), respectively).
The smaller Hall crosses are prepared by electron beam lithography. A thin
gate (thickness from 30 - 60 nm) is deposited onto the Hall bars, on top of
which the arrays are directly grown. This gate screens any potential fluctuation
due to the pressure of the particles. In addition, it provides good conductivity
for tunneling and can be used for adjusting the carrier concentration in the Hall
measurements. Either gold or soft-magnetic permalloy (80 (80%% Ni, 20%
20 % Fe) is
used as a gate material. The latter material provides a magnetic flux coupling
and hence, strong interactions between the particles. In addition, growth onto
niobium (a superconductor at sufficient low temperature) has been attempted.
The Hall measurements are typically conducted at temperatures between 10
and 100 K, magnetic fields up to 1.5 e
1. 5 T and for angles H from O to 90 of the
applied field with respect to the particles' long axis.
The gradiometer response is induced by the total magnetic stray field of
the particles. In order to calculate the induced Hall voltage, the z-component
(perpendicular to the plane of the 2DES) of stray field contributions, Hz, of all
particles within an array must be summed over the active area of the Hall
cross. The field profile of the particles is calculated exactly rather than using a
point-dipole approximation (Wirth et al., 1998b). Such an approach is
appropriate since the particle dimensions are of the same order of magnitude
302 S. Wirth and S. von Monlnar

Iv

8. 2 Schematic view of a Hall gradiometer with one of the Hall crosses carrying an
Figure 8.2
array. The 2DES is residing a certain distance below the surface (typically 80 nm). On top of
the Hall cross.
cross, a metallic gate film is deposited onto which the particles are directly grown.

as distances between particles and magnetometer. The particles' iron cores


are assumed to be cylindrically shaped (height h and radius R) and to have
their spontaneous polarization J s homogeneously distributed and oriented
within their volume and hence, a magnetic moment m p = nR TIR 2 hila'
hIJo l J s . The
stray field contributions of the particle's two axial planes, H~, and its radial
plane, H ~, are calculated separately by integration over magnetic surface
charges. Hence,
(8. 1)

where hi is the distance in z-direction between the axial plane under


consideration (top or bottom of the cylindrical particle) and the active Hall
layer and z d is the thickness of the layer below which the active layer is buried
underneath the substrate surface. In the case of the axial planes, the integral
can be solved

h + q + qR
H~ = JJS~i~~Se
sh;cose {h~ h~'ji-+q::+q~)2
2
+ 3- [ 2u_ arctan(.J2u-)
2C -
+-
2
H; {hi -
I
4q3y'2[ 2u_ arctan(J2u.)
8110 h~Jh~+(q+R)2 4q.J2

- 2u- arctan(Uv-(c-
-2u- U+ In ~!2u+
R))-
arctan ( u - (c - R)- u+ l-.J2u+
v l+,J2u+
l+.J2u+
+ u+ In v - u+ (c + R) ] } (82)
v-u+(c+R)
with

c = J h 7 + q2,
/111 (8.3a)
u= Jq/(cq), (8.3b)

J
vv = ) ~ (h7
(h~ + q2
q2 + R 2
2 ))+
+ qR.
;-;;. (8.3c)
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 303

H'
Hz =- Jscos( cp - Ij>q)sinef"
Jscos(1j> cpq )sinef" coslj> 'd'
cp [
cos cp 'd'[
Ij>
1 ,
] I Zd+
R 2 - 2qRcoscp
Z

2TrIJo 0 Jh7 + q2 + 2qRcoslj> h;=Zd


h'=Zd

(8.4)
where q is the radial distance between the particle's symmetry axis and the
point in the Hall cross under consideration. The angle e between particle
magnetization and its long axis is obtained from reversible magnetization
rotation (see Section 8.4. 1). For the radial plane one finds with q> being the
azimuth angles of the particles' magnetization and Ij>q cpq the corresponding
azimuth angle, the remaining integral is solved numerically. All of the
particles, contributions are then summed up over the active area of the Hall
cross and converted to a resulting Hall voltage. Such Hall curves are compared
to experimental ones to calibrate the average particle radius R. The distances
needed for the calculations are taken from AFM and MFM images or, in the
case of z d' are known from the fabrication process. The carrier concentration
is obtained from the empty Hall cross. A possible depletion at the edges of the
Hall cross is not taken into account. Note that, for the dimensions of interest
here, the calculated stray fields differ by at most 5 % compared to values for a
rod of height h but with square base planes of the same area instead of circular
planes.
The result of a stray field calculation (z-componenO applied
(z -component) for zero appl ied field
and for a plane z d == 160 nm below the particles' base is shown in Fig. 8. 3
(location of the 2DES for this sample). At this height, most of the local
magnetic features of the array are smeared out since the interparticle distance
was only about 100 nm. This allows a simple calculation of the Hall voltage
from the calculated stray field by using the properties of the Hall cross (carrier
concentration, Hall cross dimensions). In fact, a mean stray field value over
the active Hall area is expected to be a reasonable approximation even when

1.0
!='
f='
5-S
:2 0.5
'"
;.;:
<.::

20
Figure 8.3
8. 3 Calculated stray field (z -component) of an array of 7 x 11 particles (rows are
(z-componenO
rotated - 45' with respect to the x and y-axes shown). The z-component of the stray
field is shown for a plane parallel to but 160 nm below the particles' base plane, i. e. for
the plane of the 2DES. At this distance, the local stray field is varying only little.
304 S. Wirth and S. von Monlnar

calculating the Hall response of more inhomogeneous stray fields due to larger
particle distances (Li et al. , 1997b). This is caused by the usage of the Hall
crosses within the ballistic regime. Note that in the diffusive regime the active
area of the Hall cross can be larger than in the ballistic regime due to a
possible spread of the transport current into the voltage probes (Peeters and
De Boeck, 1999).
Stray field calculations are also performed for a height above the
particles' top corresponding to the MFM scan height. By comparison with
MFM data, the relative magnetization of the particles within the identical array
is inferred.

8.3.3 Optimized Hall Measurements for Particle Arrays

In the case of our STM-grown particles the magnetic volume (3 x 10- 24 - 5 x


1O- 23 m3 iron per particle) is orders of magnitudes smaller than in other Hall
experiments (Geim et al. , 1997a; Monzon et al. , 1999; Meier et al. ,2000).
, 2000).
Therefore, optimized sensitivity of our Hall gradiometer is crucial for a reliable
quantitative magnetometry of our arrays (Wirth et al. , 2000a). Specifically,
the coupling of the particles' stray field into the Hall cross must be optimized.
To this end, the routine developed above is applied for calculating the Hall
response of a typical test array grown onto a Hall cross of different size and all
other parameters remaining unchanged. (It should be kept in mind that the
properties most easily controlled in our fabrication process are the relative
size and location of the array with respect to the Hall cross; we do not intend
to discuss the properties of different 2DES materials at this point.)
point. ) As shown
in Fig. 8.4 (circles), the arrays' stray field is most effectively converted into
Palticle/2DES separation (nm)
Particle/2DES
oa 100 200

.....
'

I..
, .~
.,;;.\1
..i;7 .
ao 23456 7
Hall cross size / Particle array area

Figure 8. 4 Comparison of Hall voltages calculated by assuming Hall crosses of different


sizes (circles, lower axis). Array and Hall cross were either assumed to be aligned with
their centers (closed circles) or with one of their corners (open circle). All triangles
represent experimental results. In addition, the calculated Hall voltage for different
separations between particles and 2DES is shown (crosses, upper axis).
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 305

Hall voltage if the Hall cross size does not exceed the array size. In this case,
the electrons in the active area of the Hall cross are optimally influenced by the
stray field and the stray field causes the highest potential over the complete
width of the voltage leg. Compared to arrays fabricated earlier (Fig. 8. 4,
open triangles) one should be able to increase the sensitivity by an order of
magnitude just by matching the sizes of the array and Hall cross. If the Hall
cross is smaller than the array, only that corresponding portion of the array
causes a Hall voltage. The total stray field at the center of an array is,
however, sl ightly smaller than at an edge or even at a corner of the array.
This effect, which has been noted before (Gider et al., 1996), is due to
fluxlines near the center particles which form closed lines within the active
area of the Hall cross and do not contribute to the Hall voltage. Hence, the
Hall response is reduced at the array's center (closed circles). In contrast,
for a small Hall cross located with its edge underneath the corresponding edge
of the array, the Hall response is slightly increased (open circle).
Also shown in Fig. 8. 4 is the dependence of the Hall response on the
parameter Z d. This distance between the particles' bottom and the 2DES can
be influenced to some extent by the thickness of the gate material. Its lower
bound, however, is determined by the distance of the 2DES from the substrate
surface as given by the molecular beam epitaxy (MBE) fabrication process.
From our calculations, Z d is expected to have only minor influence on the Hall
response if kept within reasonable limits.
The most intriguing result from the calculations presented in Fig 8. 4 is
certainly that the expected Hall voltage depends mainly on the relative size of
the array and Hall cross but not substantially on their absolute size (of course,
the Hall cross should be kept in the ballistic regime). This result points toward
the prospect that arrays of any number of particles, even one particle, might
be measurable if grown onto a Hall cross of appropriate size.
An array fabricated to optimize Hall sensitivity by size match of array and
Hall cross is shown in Fig. 8. 5. The array consists of 420 particles with
interparticle distance of about 180 nm. It is grown onto a Hall cross of about
3.2 IJm x 2.8 IJm in size that can clearly be recognized in Fig. 8. 5. The
dimensions of the particles are chosen to coincide with those of earlier growths
to allow for comparison of the measured Hall voltages. As shown in Fig. 8.4, 8. 4,
the Hall voltages measured on optimized arrays (closed triangles) exceed
those of earlier growths (open triangles) by an order of magnitude (after
adjusting for changed experimental conditions, e. g. drive current and carrier
concentration of the 2DES). (This improvement can also be seen in
Fig.8.12).
Fig. 8.12). We will show later that this gain in Hall voltage and hence,
resolution in the magnetization curves allows for a detailed study of different
magnetization processes in these particles.
The maximum Hall voltage for the array shown in Fig. 8. 5 is measured to
be .....,
- 0.62 IJV (note that the Hall voltage depends on the magnetization
configuration within the array). This is somewhat lower than our calculated
306 S. Wirth and S. von Monlnar

II~m
Jlrn

Figure 8.5 Array of 420 iron particles grown onto a Hall cross of matching size for
optimized Hall sensitivity. The etched Hall cross of about 3.2 ~m x 2.
3. 2 IJm ~m is clearly
2.88 IJm
visible.

prediction (Fig. 8. 4). Most likely, likely. this small reduction is caused by growth
imperfections of the particles. The noise of the Hall voltage is measured to be
0.04 IN IJV 1Hz 1/2 at 30 K and zero field and about 0.07O. 07 IJV 1Hz 1/2 at 100 K and
applied
appl ied field of 1. 1.00 T. Using these data we can now calculate the required size
of the Hall cross for arrays of any size. It should be noted that the Hall cross is
fabricated first. Hence, an optimized fabrication process involves the following
steps: CD Deciding on the dimensions of the particle and their number and
arrangement within the array and calculating the required Hall cross size;
(2)Fabrication of the Hall cross; @ Deposition of the gate material (non-
magnetic, magnetic or superconducting) and lift-off; @)Growth @Growth of the particles
onto the Hall cross at its exact location. This last step requires that the
particles to be investigated can be grown at a predefined position with high
purpose, 8TM assisted CVD appears to be a
spatial resolution. For this purpose.
perfect tool: the 8TM is first used in imaging mode to determine the position of
the Hall cross and then the growth is initiated at the defined position. Another
fabrication process that allows for precise positioning of the particles is
electron beam lithography (Geim et al., ai., 1997b; Monzon et al.. al., 1999), so
far, however, with much larger particle dimensions.
far.
An array consisting of 16 particles has been grown onto a Hall cross of
about 1 IJm x 1 IJm in size (Fig. 8. 6). Hall measurements on this array have
confirmed our calculations. Using these results, we estimate a Hall voltage of
..... 0.24 IJV for a single particle of 10 nm in diameter grown onto a 400 nm x
'"
400 nm Hall cross if the same experimental conditions are used as above. This
voltage would exceed the highest noise level by a factor of 5. Hence. Hence, we
expect to be able to measure arrays of any number of particles, from a single
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 307

particle to a few particles and up to several hundred particles, by growing


them onto Hall crosses of appropriate size. The motivation here is to gain
insight into the development of magnetic properties and interactions between
particles.

500 11m
50011111

Figure 8.6
8. 6 SEM image of an array consisting of 16 particles grown onto a Hall cross of
~m x 1 IJm
about 1 IJm ~m in size. The offset of the array toward one edge of the Hall cross is
expected to further optimize the Hall sensitivity as discussed in the text.

One possible issue in making ever smaller Hall crosses is the depletion of
carriers close to the etched contours in the GaAs/GaAIAs 2DES. This effect,
which may extend several tens of nm from the edges of the Hall cross, reduces
the active size of the Hall cross (Bending, 1999). Hence, the choice of
materials for the 2DES should not only be made with respect to carrier
concentration and mobility but also depletion effects and temperature range
over which the 2DES can be used.

8.3.4 Variable Field Magnetic Force Microscopy


( MFM) in Nanoparticles
(MFM)
Magnetic Force Microscopy (MFM) combines high spatial resolution and a
relatively high sensitivity with ease of use. Since MFM is a form of SPM that
utilizes sharp-pointed tips, the corresponding equipment is often very
versatile. Such devices have become very popular and MFM is often the
method of choice for investigating magnetization configurations on a micron
length scale and smaller. We use a commercially available Dimension 3000
SPM made by Digital Instruments with home-made magnetic field and
temperature stages.
In MFM, the signal is detected while a magnetic tip is scanned over the
area of interest. The tip is kept in an oscillating motion perpendicular to and at
308 S. Wirth and S. von Monlnar

a well defined height above the investigated surface. If an AFM line scan is
performed beforehand, this information can be used to keep the tip-surface
distance constant during the subsequent MFM scan; a method that allows for
some discrimination between topographical and magnetic influences. A
magnetic interaction between tip and sample results in a change of ampl itude,
frequency and phase of the tip's oscillating motion which is detected by laser
light reflected from its backside. We track the phase shift of the tip since it is
least influenced by other than magnetic effects (Porthun et al. , 1998).
The magnetic tips are usually fabricated by coating a silicon tip (e. g. , an
AFM tip) with a magnetic (e. g., CoCr) film (Babcock et al., 1994;
Leinenbach et ai.,al., 1999). The magnetic tip end radi radiii of about 50 nm
influences substantially the lateral resolution of MFM. MFM tips have also been
fabricated by electron (ROhrig
(Ruhrig et al.,
aI., 1996b) and ion beams (Folks et al. ,
2000) .
Even though MFM images are often very intuitive (Rave et al., 1998;
Proksch, 1999) they are difficult to interpret quantitatively. An exact
quantification requires detailed information on the topography and magnetic
(Gbddenhenrich et ai.,
configuration of the tip. Current strips (G6ddenhenrich al., 1990),
microfabricated current loops (Lohau et ai., al., 1999) and electron holography
(Matteucci et ai.,al., 1993; Streblechenko et al., 1996) are used for tip
calibration. An approximate description of tip magnetization in terms of
magnetic monopols and dipoles has been given by van Schendel et al.
(2000). A more severe issue - at least if soft-magnetic samples are to be
investigated - is the interaction between tip and sample. The tip's magnetic
stray field (Thiaville et al.,
aI., 1997) may locally disturb the magnetization
configuration inside the sample and complicate small (and zero) field
investigations.
In the case of our particle arrays, MFM is util ized to extend the
temperature range over which the particle's magnetization reversal is
investigated up to room temperature. Moreover, AFM/MFM provides data on
the local arrangement of the dots with respect to each other and the Hall cross
as well as the relative magnetization of the dots. The latter can be extracted
as long as all parameters involved in MFM are kept constant during the scan.
A home-made magnetic field stage (based on an earl ier design by F. J
Cadieu et al. (1997)) is used to apply fields of up to O. 3 T perpendicular to
the plane of MFM scan (i. e., perpendicular to the substrate and hence,
parallel to the long axis of our particles). The field is produced by a
cylindrically shaped NdFeB permanent magnet (Vacodym 510HR by
Vacuumschmelze Hanau GmbH, remanence of 1.41 T) magnetized
perpendicular to its rotational axis. Iron yokes below and above the permanent
magnet are used to collect and homogenize the magnetic flux to some extent
(Fig. 8.7).
8. 7). The field emanating from the top of the upper yoke can be changed
continuously by rotating the permanent magnet cylinder. The actual applied
field, H app' is measured by a Hall probe very close to the sample. This stage
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 309

is only about 2 cm in height and fits into the 03000


D3000 SPM. Somewhat higher
appl ied fields can be obtained from a stage with the field parallel to the scan
area because of a better flux closure through the air gap between the yokes in
this configuration (Proksch et ai., 1995). A different approach was reported
by using a pair of external Helmholtz coils (Porthun et ai., 1998; Kleiber et
al.,1998).
al. , 1998).

/.2:::":::;7'/,-
-----7'7;L-\--- Hall sensor

~L-
7 4L_
- - -_--+-_
-+ - - Sample

Permanent magnet
(can be rotated)

-----,~--Yokes
-----c~--Yokes

Figure 8. 7 Schematic view of the magnetic field stage as used in the MFM. A magnetic
field of up to 0.3
O. 3 T is applied
appl ied perpendicular to the scan area of the MFM and the substrate.
The field can be changed by rotating the permanent magnet (shown is the maximum field
position) .

To study the magnetization reversal in our arrays of nanoparticles, its


magnetization is initially saturated by applying the maximum field parallel to
the particles' long axes. H app is then applied in opposite direction and
switched off before taking the MFM image. The magnitude of H app is increased
stepwise and images are taken until complete reversal of the array has
occurred. This procedure ensures that the particle magnetization is reversed
only by the external field and not by a superposition of H app and the tip stray
field. Moreover, it becomes obvious from the images if the magnetization
reversal is caused solely by the tip stray field, which is estimated to be about
40 mT (Tomlinson and Hill, 1996) for the conditions used.
A sequence of MFM images taken from the same part of an array for
increasing I H app I but otherwise identical conditions is shown in Fig. 8. 8 (for
(tor
comparison, a corresponding AFM image is also shown). The zero field image
was taken after saturating sample and MFM tip. After applying a field pulse of
- 45 mT in the opposite direction two of the particles within the investigated
area reversed their magnetization as indicated by the bright spots (in our
setup, the magnetization appears dark if parallel to the tip magnetization and
vice versa). Moreover, the contrast becomes less pronounced due to a more
inhomogeneous magnetization configuration within the MFM tip (Babcock
et ai.,
al. , 1996). The latter is even more obvious for H app =
Happ - 50 mT, an image
taken just before reversal of the tip's magnetization (not all images taken are
shown). This tip reversal caused the majority of the particles to appear bright
310 S. Wirth and S. von Monlnar

in the - 60 mT image. Note the drastically increased contrast in this image. It


is speculated that the magnetization at the very end of the MFM tip reversed
first. This reduced the tip's total magnetic moment and hence, the tip-particle
interaction. After reversing the complete volume of the tip's magnetic coating
the maximum contrast is recovered. The field at which the tip reverses its
magnetization depends strongly on the type of tip used. With decreasing H app ,'
the particles progressively reverse their magnetization. This process is
completed at around 90 mT. A mean particle switching field of H sw = 65 mT at =

500 nm

(a)

OmT
~.

.,.
. ,......
-. ..
-45 mT
-4SmT

c
, .::.j:.;
:~

-50mT
-SOmT
. ..
~.I I.- .;1...~
---.
--60 mT
-60mT
""""' .....

V
i
i
-70mT
(b)
" -90mT

Figure 8.8 AFM image (a) and a sequence of MFM images (b)Cb) taken from the same part
of an array. Reversed fields of increasing magnitude are applied. The magnetization
reversal within the MFM tip as well as the individual particles is obvious. See text for
discussion.

room temperature is inferred. The strongly local character of the tip's stray
field can be used to reverse the magnetization of a predefined particle. This
process corresponds to bit writing (Kong et ai.,
aI., 1997; Kleiber et al., 1998)
and is demonstrated in Fig. 8.
8.99 (Field, 1998). Here, the same array as in
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 311

Fig. 8.8 is used. After an initial MFM image was taken (a) the tip was placed
over the target particle in the upper right corner and a field pulse of - 20 mT
was applied. Tip stray field and Happ
H app were oriented opposite to the particle's
magnetization. Neither of these fields alone could reverse the particle
magnetization. In combination, however, they exceeded the particle's
switching field (b) and caused it's magnetization reversal. The field values
agree well with those given above.

(a) (b)

Figure 8. 9 Two MFM images showing the very same particles (same Csame array as shown in
Fig.8.8).
Fig. 8. 8). The magnetization of the top right particle is reversed in (b)
Cb) with respect to its
original erientation Ca) Cb) by the combined action of the tip stray field and of an external
(a) (b)
field pulse C-
(- 20 mT). All other particles keep their original magnetization orientation.

8. 4 Magnetization Processes in Small Particles

8. 4. 1
8.4.1 Reversible Rotation of Magnetization
In the following, the experimental investigation of basic magnetization
processes, reversible rotation and magnetization reversal, will be described.
Even though these processes are common to all ferromagnetic materials, we
will concentrate on the specifics of small, single domain sized particles.
Our 8TM-assisted CVD results in pillar-shaped particles. These particles
have aspect ratios from 4 : 1 to 20 : 1 that induce shape anisotropy. A shape
anisotropy constant,
=
K s = (N~ - Nil) JU2J,Jo
JU2!Jo (8.5)

as high as 750 kJ/m 3 can be expected if the particles consist of a single grain.
Here, Nil and N ~ are the demagnetization factors parallel and perpendicular
N..L
to the particles' long axis respectively. Magnetocrystalline
Magnetocrystall ine anisotropy in iron
is much smaller (53 kJ/
kJ/mm3 ) (according to Kneller, 1962) and is neglected.
Hence, the particle's long axis forms an easy magnetization direction (EMD)
(EMD)..
If the field is applied parallel to the EMD (8 H = 0) reversal of the
312 S. Wirth and S. von Monlnar

particle's magnetization is observed, whereas for a field appl ied


perpendicular to the EMD (e H = 90') reversible rotation takes place. For any
other acute angle 0'<
0' < e<
H
H 90' both processes will contribute to the particle
magnetization change. To study the reversible rotation we therefore
e
concentrate on H """""'" 90'. e
In Fig. 8. lOa, HH = 90' is obeyed exactly. We start
with a completely saturated sample which results in the maximum Hall voltage
(marked by I ). For increasing field ( II IT ) the magnetization rotates toward
the field direction, the stray field in the z-direction decreases and a lower Hall
response is measured. For sufficiently high field H app H A (the anisotropy >
field) the magnetization is aligned parallel to the field (ill), a configuration in
which no stray field in the z-direction is generated and hence, no Hall
response is measured. If the field is now decreased below H a the
magnetizations can rotate toward either easy direction. For HH = 90', both e
orientations are equally probable and no net magnetization in the z-direction is
obtained. This symmetry is supported by thermally activated switching
(Section 8.5) at fields just below H H a'
a In zero field (IV) the sample is field-
demagnetized with the magnetization of half the particles pointing in one
direction of the EMD, the other half in the opposite direction. The sample could be
remagnetized by applying a field at HH """ e
90'. The fact that an array can be field-
""'" 90.

demagnetized strongly supports the applicability of the magnetization rotation model


on which our analyses will be based (Fig 8. lOb). lOb) .
~

>::t
~ 0.05 T=60K, e=90.0'
T=60K,8=90.0'
0.05

~v
:::J
<l

ioD 0.00
Zi 0.00
__.-'' ' ' ' 1IIt1
~ 1
IV
\",1 tWI,Il~.
III

.~~
)H
~ .t,~'
"0 H
> -D05
"0 -0.05
~
-;; . z ---J I~~
1~~J{,ffl
-D.lO!--o:-_ _--::-,-----_~~,-----_~-'.,,-':-II_ _----::e
:c
:r: -0.10
-0.8
-D.8 -0.4
-D.4 0.0 0.4 0.8
/loH
Applied field lloH (T)
(a)

T=30 K, 8 =89.3'

... ~~l~~.~~,
0.1
>
2- ~". .,~
:::J 0.0
0.0 ,.~
"'.,.~ ~"
<l

-0.1 r' ~~~ .

-D.6 -D.4 -D.2 0.0 0.2 0.4 0.6


lloH (T)
(b)

Figure 8. 10 Magnetization curves of different arrays measured for applied field perpendicular
to the particles' long axes (EMD). Due to the rotation of the particles' magnetization toward
the field direction, I.i. e., away from the sensitive direction of the 2DES, the Hall voltage
diminishes for Increasing [e HH =
increasing field. After applying a field exactly perpendicular to the EMD [8
90', curve (a) ] the sample is Is field-demagnetized (see text). This symmetry effect disappears
sl ightly deviating field angle [e HH = 89.3',
for a slightly 89. 3', curve (b)].
(b).:.
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 313

The symmetry can be broken by a small deviation from the perpendicular


field orientation as shown in Fig. 8. lOb for 8 HH = 89.3.
89. 3. The resulting minute
field component along the EMD forces all particle magnetizations to rotate in
the same direction if H app
app is decreased below H a'
a . The maximum Hall response
is recovered at zero field. Moreover, the Hall voltage does not reach zero for
the highest fields applied. This is caused by the small remaining magnetization
in the z-direction due to the field misorientation.
In the following, the reversible magnetization rotation is analyzed further.
In our nanoparticles, a uniform magnetization orientation is assumed;
deviations certainly cover only small volume fractions (e. g. , less than 1% at
the particles' end caps as found by numerical simulations, Section 8. 4. 5) .
Hence, exchange energy contributions to the particles' free energy density f
vanish. Taking into account Zeeman and anisotropy energy densities yields
(Stoner and Wohlfarth, 1948)
f = K ssin 2 8 - JsHcos(8 H - 8). (8.6)
The equilibrium orientation of the particle magnetization 8 (H, H ) is found by
(H , 8 H)
minimizing f with respect to 8 and is subsequently used in the Hall voltage
calculation as outl
outlined
ined in Section 8. 3. 2. The corresponding results are
included in Figs. 8.10 as white lines. In these calculations, K s is considered
as a fitting parameter in order to enable comparison with values expected from
the particles' dimensions. The mean values of K ss as determined from such fits
are K ss = (0. 61 0 . 04) MJ/m3 and K ss = (0. 40 0 . 04) MJ/m3 for the particles
(0.61
in Figs.8.
Figs. 8. lOa and 8. lOb, respectively. Note that the kink in the magnetization
curve at Ha = 2K s /JJ s is only observed for 8 H =90
= 2Kj = 90 (Fig.8.10a).
(Fig. 8. lOa). The more
the direction of the applied field deviates from the exact perpendicular
orientation, the less pronounced is the curvature observed in the magnetization
curves at around H a' The overall agreement between measured and calculated
curves (Fig. 8.10) supports the applicability of the reversible rotation model.
Moreover, K ss is found (Wi (Wirth
rth et aI., 1998b) to be temperature independent.
independent
for the investigated temperature range 10 - 100 K as expected from
Eq.(8.5).
Structural inhomogeneities of the individual particles lead to a distribution
of K ss within an array. Such a distribution of K ss is now taken into consideration
in calculating f from Eq. (8. 6) and, consequently, the total magnetization of
the array. The analysis of reversible rotation as outlined above can even be
applied for multigrain or (magnetostatically) interacting particles (Wirth,
1998). In contrast, irreversible processes can only be analyzed if the exact
magnetization reversal mode is known [one example is uniform magnetization
reversal in extremely small particles; (Bonet et al., 1999) J. However, the
aI., 1999)].
switching field H sw ((88 H ) approaches the value predicted for coherent
magnetization reversal in non-interacting small (single domain) particles
independent of the reversal mode if 8 H approaches 90
H sw = H a (sin 2/3 8 H + COS2/3 8 H )-3/2. (8.7)
314 S. Wirth and S. von Monlnar
MonlncH

Hence, irreversible processes can also be evaluated if BHH is close to 90". 90. e
Equation(8.7) appl ied to the case BHH =
Equation (8.7) may not be applied == 90 e
90" because of possible
field demagnetizing effects as discussed above. With this information, information.
hysteresis curves taken in the first/third and second/fourth quadrant can be
fitted simultaneously (Wirth et alal.., 2001). In the former,
former. only reversible,
reversible. and
in the latter both reversible and irreversible processes are taken into account.
The result of such a fitting procedure ([J,(D. 0) is presented in Fig. 8. 8 11 for the
17 nm particles and BH = e
== 86"
86 along with the measured magnetization curve
( x ). The relative number of particles switching within a certain field interval
(line,
(line. right axis) is determined from the additional change in magnetization in
the fourth quadrant compared to the fifirst. rst. In this calculation,
calculation. particle
magnetizations already reversed at lower fields are taken into account and the
equilibrium orientation for non-switched and switched magnetizations are
obtained by Eq. (8.
(8 6). Note that the anisotropy (K s) -distribution influences
both the reversible [via Eq. (8.6)J
(8.6) J as well as the irreversible processes [via
(8. 7)J as a result of the constrain BHH = 90.
Eq. (8.7)J 90". Hence, e
Hence. the line in Fig.8.11
Fig. 8.11
corresponds to an anisotropy distribution.

.L
0.3 T=30 K 2 .~
881/=86'
H =86
'0
0.2
6:
I 2
~
>
0.1
.~c:
3:::l. r-----",,,..---+--'------:l~-____l0 <
~
<l
-0.1

-0.2
-D.2

-o.3,--~-::'-c-~--=-----"L,------::-,-:------:::'-c-~---,-------J
-D.3'----::"-:c-::'-c-~--=---::-'-:,-----_:_'_:_-----:~~--,-------J
-0.6
-D.6 -0.4 -0.2 0.0 0.2 0.4 0.6
lloH (T)

Figure 8. 11 x ) of a particle array (mean diameter 17 nm) for field


Hysteresis curve ( X)
applied at 86' with respect to the particles' EMD. Fits account for reversible (first and
third quadrant( D) and a combination of reversible and irreversibie
irreversible (second
Csecond and fourth
quadrant( 0) magnetization processes. The" difference" is then related to irreversible
quadrantC
contributions which allows an evaluation of the anisotropy distribution (line, right axis) .

The unusually broad anisotropy distribution is the broadest we found for


our particle arrays so far. As a general trend, the larger the particle diameter,
diameter.
the broader the anisotropy distribution. This suggests an interpretation that the
different peaks in anisotropy (Fig. 8.11)
8. 11) may be related to particles consisting
of a different number of grains. From Eq. (8. 5) the hardest particles seem to
consist of a single grain. Particles consisting of two or more grains are
magnetically softer. Typical values of the Hsw-distribution for our arrays are
about 50 mT (Fig. 8. 12b, Table 8. 8 1). Along with the observed values K ss we
believe
bel ieve that most of the smaller particles consist of a single grain.
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 315

Microfabrication techniques are expected to allow for control of the


microstructure within the particles. However, very large variations in the
particle switching behavior are reported due to structural inhomogeneities
(Kleiber et al., 1998; Haginoya et aI.,
al., 1999; Wegrowe et al., 1999; Ross
et al. , 2000). Thus 8TM-assisted CVD appears to offer an advantage.

170m
0.5 0
811 =0
~.

>'
>::L
:::l. 0.0
::..
<l
<I
-0.5._~
-0.5

/loHapp(T)
f./c/lapp (T)
(a)

-0.4 -0.2 0.0 0.2 0.4


f./l)H
J1DHappapp (T)
(b)

Figure 8. 12 Magnetization curves for field applied parallel to the particles' EMD. (a)
Same array as in Fig. 8. 11. The unusually broad anisotropy distribution of this array is
reflected by the" bump" in the magnetization reversal. (b) Magnetization curves of an
the "bump"
array of 14 nm particles showing a significantly smaller switching field distribution. The
higher Hall response of array (a) is achieved by an optimized geometry as described in
Section 3. 3.

Table 8. 1 Properties involved in magnetization reversal of particles of different diameter.


(T=
H sw (T = 0) are the switching field values extrapolated from H sw ( n to zero temperature.
(T)
llH sw = (H ~ift
The distribution of switching fields is characterized by fJ.H ~.;;) 12. V" is the
~iJt - H ~~)
mean critical volume as derived from Eq. (8. 12). H ~ and H; are the calculated coercivities
for long and short times, respectively, at room temperature. For a discussion of the values,
specifically on the 19 nm-particles, see also section 8.5.3.
8.5. 3.
Mean diameter, d(nm) 9 10 14 17 19
HswCT=O) (mT)
110 Hsw(T=O) 242 288 223 165 135
110 fJ.H
llH sw (30 K) (mT) 45 40 40 70 60
3
v"
Vcr (nm ) 210 260 470 720 1200
vcrl d'2 (nm)
veri d 2.6
26 2.6 2.4 25
2.5 3.3
110 H~ (mT) 2 21 48
110 H; (mT) 170 232 192 145 132
316 S. Wirth and S. von Monlnar

8.4.2 Models for Magnetization Reversal without


Thermal Activation
Magnetic materials in general can show very complex magnetization reversal
behavior. This is mainly due to a competition between long-range
magnetostatic and short-range exchange interaction which leads to the
formation of domain walls. If the particles, however, become small enough,
they consist statically of only one magnetization domain (single domain
particles, usually a few 100 nm in diameter). Yet, as long as one particle
dimension is still larger than about the domain wall width of the material
(usually a few 10 nm), the magnetization reversal is non-uniform as we will
illustrate in the following sections. Only for particles as small as 20 nm in all
dimensions uniform magnetization reversal is observed as, e. g. in the
excellent experiments by Bonet et al., (1999). Only in this case can the
simple Stoner-Wohlfarth model (or its three-dimensional extension) be applied
reliably. This model predicts switching fields as given by Eq. (8.7) for uniaxial
anisotropy that can be described by a single constant.
For cylindrically shaped particles of infinite length the inhomogeneous
reversal modes buckling and curling were introduced (Frei et ai., aI., 1957;
Aharoni, 1966). These modes exhibit considerably reduced switching fields
e
for eH == 0 if compared to homogeneous reversal. The nucleation field (which
here equals H sw ) in the case of curling is given by

H
/-IoH
sw
-- Js p(l
p(l-- p) (8.8)
J.1o sw - - 2 J p + (1
22 - 2p)COS 2e
2p)COS2eHH

~x )) 2,
where p = 1.08 ( ARx 2 , A
A.exx = (/-10 A) 1/2 / J s is the exchange length and A the
(J.1o

exchange constant. Hence, a d - 2 dependence of H sw ( T = 0) is expected.


Note that buckling may occur for a small range of diameters between those for
homogeneous rotation and curl
curling
ing (Frei et al. , 1957).
Results of numerical simulations on magnetization reversal are discussed
in Section 8. 4. 5. Magnetization reversal in magnetically inhomogeneous
nanostructures has been considered by Skomski et al. (1999, 2000).

8.4.3
8. 4. 3 Experimental Results on Magnetization Reversal
As already pointed out, the magnetization reversal is best observed with field
appl ied parallel to the particles' EMD. Corresponding results for an array of
17 nm particles (same array as used in Fig. 8.11)
8. 11) and of 14 nm particles are
shown in Fig. 8. 12a, b, respectively. For the latter, a mean value of the
switching fie'ld, H sw ', of 280 mT can IlH sw of
c;an be extracted with a distribution b.H
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 317

about 45 mT. Magnetization curves as shown in Fig. 8. 12 are used to


determ ine the switching field dependence on temperature, field angle 8eHHand
determine and
particle diameter d.
For fields applied parallel to the EMD (8 (e H = = 0), the temperature
dependence of H sw for the different d is used to estimate "zero" temperature
switching fields H sw (T = 0), i. e., Hsw-values without influence of thermal
activation. The numbers obtained (Table 8. 1) are much smaller than
predicted by uniform magnetization reversal and hence, the particle
magnetizations reverse by inhomogeneous modes. One likely mode is curling
as discussed above. The expected d- 2 -dependence of H sw ( T = 0) is
approximately observed for the bigger particles d ? ~ 10 nm. From Eq. (8. 8)
we estimate Aex = = 3.1
3. 1 nm and A --- 36 pJ/m in agreement with values by
Scheinfein et al. (1991). For the smallest particles, the reversal mode could
be more influenced by inhomogeneities or may be of buckling type (Frei
et ai.,
aI., 1957; Aharoni, 1996).
In addition to H sw ((T
T = 0), the angular dependence of H sw (e ( 8 H)
H ) of the
14 nm particles at 10K is compared to values expected from Eq. (8. 8). As
shown in Fig. 8. 18a, agreement is found for e :::;;; 30 (using again Aex
8 HH :S;;;30 ex =

3. 1 nm). For e 8 HH > 30 the measured H sw (e H )-values


(8 H) -values follow results calculated
for a uniform mode since it requires lower switching fields than curling. Such
transitional behavior from curling to homogenous reversal was predicted (Frei
et ai.,
al., 1957; Aharoni, 1996) and experimentally also found for Ni-wires
(Lederman et al., 1994; Wernsdorfer et al., 1996). Note that curl ing and
micromagnetic simulations give similar Hsw-values for the particle dimensions
R = (23)
(2---3) Aex considered here (Schabes, 1991; Suhl and Bertram, 1997).
Numerical simulations (Rave et ai.,
al. , 1998) indicated that the exact shape
of extremely small particles (like sharp corners) does not drastically influence
their switching due to the dominance of exchange interactions. Hence,
differences in d for different arrays appear more relevant than different
morphologies. Switching at higher temperatures (Fig. 8. 18) is discussed
within the framework of thermal activation (Section 8.5.3) .

8. 4. 4
8.4.4 Interaction Effects
For our typical arrays the stray field emanating from one particle to a
neighboring one is below 1 mT. Therefore, interaction effects cannot be
observed reliably. To enhance interactions we grew several arrays onto thin
permalloy films. Permalloy is a soft-magnetic material and is expected to
"guide" the particles' magnetic flux. By interactions exaggerated in such a
manner we mimick the magnetization behavior of an ever smaller and more
dense magnetic storage media (Wirth and von Molnar, 2000b). Moreover,
heterostructures involving magnetically soft and hard components can be
studied.
318 S. Wirth and S. von Monlnar

60 nm thin films of permalloy (80% Ni, 20% Fe) are deposited onto Hall
bars (these films replace the normal Au gate, see Section 8. 3. 2). Iron
particle arrays were then grown directly onto the permalloy. Examples for
different particle arrangements of 13 nm particles (height 100 nm) have
already been shown in Fig. 8. 1 of Section 8.2.2.
8. 2. 2.
SQUID magnetization was measured for temperatures 5 - 300 K and
varying sample orientations. These measurements reveal only the overall
permalloy magnetization behavior (due to the several orders of magnitude
smaller particle volume, the Fe magnetization cannot be detected by SQUID).
SQUID) .
A preferred orientation of the permalloy magnetization in the film plane is
confirmed. For fields applied
appl ied perpendicularly to the film, the magnetization
saturates at about 1. 1 T agreeing, with the permalloy saturation polarization,
J s=1.05T.
Striking evidence for interparticle interactions comes from MFM
investigations. Examples of neighboring particles with antiparallel
anti parallel
magnetization orientation are observed to switch their magnetization
simultaneously under the influence of the tip stray field while being scanned.
This tip induced magnetization reversal can be reversed by reversing the MFM
slow scan direction and is repeatable.
Considering the particles only, an anti parallel magnetization orientation
for adjacent particles is energetically favorable (checkerboard configuration)
configuration)..
For the permalloy, on the other hand, no major change in magnetization
orientation is expected on the length scale of interest here. Due to flux
coupl ing, these two effects are competing in our heterostructures. As a result,
coupling,
we find lines of particles with parallel magnetization orientation and adjacent
lines having opposite magnetization orientations (Fig. 8. 13). These
configurations are relatively stable (for the heterostructure of Fig. 8. 13 in
fields from - 20 to - 50 mT opposite to the initial saturation direction of the
particles). The lines are not perfectly formed since particle magnetizations
already reversed at smaller negative fields have to be accommodated. As a
consequence, the magnetization within the permalloy film is modulated on a
scale comparable to the particle's distance. This scale is much smaller than
observed in micro-patterned permalloy films without additional nanomagnets
(Runge et al.,
aI., 1996; Gomez et aI.,al., 1999). Note that these Iines are not
observed for particle arrays grown onto gold, i. e., for non-interacting
particles.
Hall measurements conducted for fields parallel to the permalloy film (8(B H
= 90) show a jump at very low fields related to the in-plane magnetization
reversal of the permalloy (2 - 4 mT in agreement with SQUID measurements)
measurements)..
For fields higher than 30 mT, the permalloy magnetization is not significantly
modified by the particles' stray field and, hence, only Hall signals related to
the homogeneous rotation of the particles' magnetization are observed. This,
again, can be used to determine K sand d. As an example, K s= 0.46 MJ/ m3
is determined for the 13 nm particle heterostructure, in agreement with values
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 319

Figure 8. 13 MFM image of a particle/permalloy heterostructure taken after applying a


reversing field of 50 mT. The particles' iron core diameter is 16 nm. Lines of alternating
magnetization orientation can be recognized.

obtained for similar particles grown onto gold. The mean particle diameter
determined magnetically, d ~ 13 nm, agrees reasonably with the one
expected from the growth conditions (12 nm).
nm) .
For fields applied perpendicular to the permalloy film (8 H = = 0)
O~) the
measured Hall voltages of the particle-permalloy heterostructures show a
complex behavior since magnetization reversal of the particles as well as
modifications in the permalloy magnetization are observed. These effects can
be separated by comparison to results obtained from samples containing only
the permalloy without the array. The Hall voltages for the heterostructures are
at least one order of magnitude higher than for the iron particles alone,
alone. thus
demonstrating the interaction-induced permalloy response to the particles'
magnetization reversal. Results of Hall measurements for heterostructures with
13 nm particles in square (.) and triangular (C.,)
.. ) arrangement are summarized
in Fig. 8. 14. Here.
Here, H sw ( T) denotes the field at which half the particles
reversed their magnetization. The values for the different arrangements do not
show significant differences and are comparable to those found for non-
qualitative
interacting particles. These results indicate that no qual itative change of the
magnetization reversal due to interactions occurs. Rather, Rather. a mere
superposition of applied field and interaction effects can be assumed for
particles as small as those investigated here [in contrast to (Dunin-Borkowski
(Dun in-Borkowski
et al., 1999) J. The interactions broaden the switching field distribution as
al.. 1999)].
shown by the error bars in Fig. 8. 14. These bars indicate the field range within
which about 80 % of the particles reversed their magnetization. At all
temperatures investigated,
investigated. few particles reverse their magnetization at zero
field whereas fields up to - 0.5 T are needed to reverse the magnetization of
the entire array. 1:J.H
IlH sw does not vary significantly with temperature since the
320 S. Wirth and S. von Monlnar

0.3

fitiHB
"0
"'<;:;v"
t;::
gp P 0.2
gf
:.c~
:.a~
t !f
B
):l
.- ::t::
::r::
~ ::
~ 0.\
on
V>

0.1 f
"''""
OJ

~
v
'" 0.0 f
0 100 200 300
T(K)

Figure 8. 14 Magnetization reversal of particles of 13 nm diameter grown in square (.)


and triangular (.)
( ... ) arrangement onto permalloy films. The markers represent the particles'
mean switching field. The interactions are indicated by error bars that represent the field
range within which 80 % of the particles reverse their magnetization.

coupling is caused by the iron particle magnetization which changes very little
at low temperatures. These findings are in qualitative agreement with
predictions (EI-Hilo et aI.,
al., 1998). The observed strong broadening of the
switching field distribution reveals the harmful influence of interactions on the
magnetic stability of individual particles.

8.4.5 Comparison to Numerical Simulations


Numerical simulations have become an increasingly recognized tool for
investigation of magnetization processes in small particles where executable
discretization can reach a length scale below the exchange length.
A boundary-element method (BEM) has been developed (Christoph
et al. , 2001) that simplifies
simpl ifies the evaluation of the magnetostatic energy, a
difficult problem encountered in using finite element methods (for FEM see,
e. g. (Fredkin and Koehler, 1990; Boerner and Bertram, 1997; Rave et al. ,
1998; Schrefl, 1999)).
1999. In our case, the stray field Hint is calculated by
dividing the magnetic structures into polyhedral elements and solving the
corresponding surface integrals exactly. The dimensions of the elements are
chosen depending on the smoothness of the magnetization distribution and the
distance from the element under consideration. This method is suitable to
investigate any magnetic hysteretic phenomena (including non-equilibrium
states) since no energy minimization is involved. The price for a more exactly
calculated stray field is a comparatively long computing time. Thermal effects
are not yet considered but could be implemented by adding a random
contribution to the local field (Brown, 1963; Brown et al. , 2000).
As a first check, non-interacting particles were analyzed (see end of
Section 8.5.1).
Magnetization reversal for an array of 5 x 5 particles onto permalloy is
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 321

simulated with dimensions as in Fig. 8. 1. The maximum discretization length


was chosen to be 3 nm which is comparable to the subdivision size requirement
(Rave et al., 1998). Because of the much larger magnetic volume of the
permalloy, only the behavior of the particle magnetization (i. e., the total
magnetic moment of the particles, m p) p ) is shown in Fig. 8. 15a. Magnetization

/.10 H app
reversal of the particles starts at fields as small as J../o app = - 0.25 T. The
reversal partially stops when a rather stable magnetization configuration is
formed which can be considered as interaction domains. The reversal of all
app = 0.6 T .
/.10 H app
interacting particles within the array is completed at J../o

2 calc. ..........................
........................... ~

::i
~
E
r:- 0
::r:

-- ~
"-
;:: -2
;::
"1;i
-4u-
-4 --'- ~ _
~loH(T)
J.1<JH(T)
(a)

::;
:>
53-;
:J
<l
<1
6

-2
meas.

~'iOK
at 30 K

-D.6
-0.6
~
-{),4
-0.4
.... J.. l

-D.2
;--.
0.0
PoH
J.1 oH(T)
(T)
(b)

Figure 8.15 Calculated particle magnetization m pp and its derivative dmp/dH (a) in
particle-permalloy heterostructures. The latter can favorably be compared to the measured
magnetization reversal in such heterostructures (b). A reversal of the particle
magnetizations within an array in two steps is obvious. This may be indicative of the line
formation as seen in MFM (Fig. 8.13).
8. 13). Smaller measured field values may be caused by
and/or
particle inhomogeneities and/ or non-zero temperature.

As a consequence of the Hall gradiometry, the experimental findings


(Fig. 8. 15b) can be compared to the derivative dmp/dHapp(Fig.
(Fig.8.15b) 8. 15a) of the
dm p /dH app (Fig.8.
numerically calculated magnetization. Magnetization reversal is observed from
- O. 15 T (with a sharp onset) up to fields of - 0.32 T at 30 K. Although an
overall qualitative
qual itative agreement between experiment and simulations is found,
the numerically obtained switching fields are shifted to higher values (by about
O. 15 T) . This may be attributed to imperfections in the physical
heterostructures that reduce their actual H sw.sw . More importantly, thermal
activation is not considered in the numerical simulations although a strong
reduction in H sw with increasing temperature in the range 5 - 300 K is
observed (Wirth and von Molnar, 2000b).
The numerical simulation reveals that the net magnetization curve of the
permalloy layer is only weakly influenced by the iron particles (also seen
322 S. Wirth and S. von Monlnar

experimentally by the small total Hall voltage in Fig. 8. 15b). The domain
patterns, however, differ strongly. This is due to the local character of the
particle stray fields. For a layer without particles, a vortex-type pattern was
obtained, similar to those seen by high-resolution Lorentz microscopy (Runge
et a!.,al., 1996). In contrast, extended linear domain walls are formed in the
layers of the particle/permalloy heterostructure. This numerical result is in
excellent qual itative agreement with our MFM observations (Fig. 8. 13) .
Furthermore, this suggests the existence of walls inside the permalloy between
particle Iines
lines with opposite magnetization orientation. The formation of these
particle lines is a direct consequence of the interaction between the particles
and the particles with the permalloy. The interaction field, Hint'
Hint, destabilizes

the first particle magnetizations to be reversed and stabilizes the remaining


magnetic structure. An intermediate, relatively stable configuration is
apparent (two discrete peaks in Fig. 8. 15) as discussed in the previous
section. Moreover, the numerically determined distribution of H sw over a field
range of about 0.2 T can be compared favorably to the experimental results
(~O. 17 T). This supports our assumption of a mere superposition of H app and Hint
(:::::::::0.
(Section 8.4.4).

8. 5 Thermally Activated Magnetization Reversal


in Small Particles

8.5.1 General Considerations

The energy necessary to overcome the barrier for magnetization reversal can
be provided by an external field and/or thermal activation (quantum tunnel
tunneling,
ing,
e. g. , discussed by Thomas et al. (1996), is not considered). In this picture,
we assume discrete directions of the local magnetization for which the free
energy features relative minima (for our particles given by the EMD). The
concept of thermally assisted magnetization reversal due to thermal
fluctuations was introduced by Neel (1949) and further developed by Brown
((1963)
1963) who assumed uniform magnetization and uniaxial anisotropy to derive a
single-energy barrier and hence, a single relaxation time. Such behavior was
only recently found in individual ferromagnetic nanoparticles (particle
dimensions up to several tens of nanometers, (Wernsdorfer et al., 1997.
This model attracted considerable theoretical research since the energy barrier
E Bs can be calculated expl
explicitly
icitly [in particular its field dependence (Victora,
1989; Coffey et al. , 1995)]. The transition rate rI ij from state i to state j is
assumed to be given by an Arrhenius form
r;j
lij = 1
To'exp(-
To exp(- EBij/k s T)
ESij/k B (8.9)
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 323

where To
TO is the attempt time ( - 10- 9 s). For the usual investigations of
(--
thermal activation, an external field is applied to make i a metastable and j a
stable state, and reverse jumps from j to i unlikely. The latter simplifies the
Master equation (Arias and Bertram, 1997) considerably and the probability of
switching from i to j yields
P ij = 1- exp (- tr i) . (8. 10)
(8.10)
This results in the often observed logarithmic time dependence of the system's
magnetization.
Experimental data, however, often disagree with the Neel-Brown theory
even for very small particles as long as one dimension exceeds the wall width
of the material (Lederman et aI.,ai., 1994; Wernsdorfer et al.,
aI., 1995b; Wirth
et al., 1999). These deviations indicate that the magnetization reversal in
elongated particles may take place inhomogeneously along the long axis via a
complex path in configuration space which does not involve a single energy
barrier. Such behavior is supported by analytical calculations (Braun, 1994).
Moreover, thermal activation is also investigated by numerical simulations.
Monte Carlo methods have been implemented (Lyberatos et al. , 1990; Nowak
et al. , 2000) or the Landau-lifshitz-Gilbert equations are integrated directly
(Chantrell and Hannay, 1997; Bertram and Peng, 1998; Brown et al. , 2000).
We also appl ied the BEM method introduced in Section 8.4.
applied 8. 4. 5 to investigate
non-interacting particles. A consistent picture emerges, which is a nucleation
type of reversal: so-called end caps are formed that are pushed into elongated
magnetic particles by the external field and/or thermal activation until the
domain walls thus produced are swept through the complete particle to reverse
its magnetization.

8.5.2 Phenomenological Model


(F ig. 8.16)
Measured switching field values (Fig. 8. 16) are at least a factor of 2 smaller
than those calculated from the experimentally obtained K s by assuming uniform
magnetization reversal. Moreover, the switching fields show a clear
temperature dependence which cannot be explained within the Stoner-
Wohlfarth model for values of K sand J s which are essentially temperature
independent in the experimental temperature range. A similar problem arises if
our data are analyzed within a model (Kurkijarvi, 1972) which allows
evaluation of thermal activation from field sweep measurements: the energy
barrier to be overcome for magnetization reversal at zero temperature is
estimated to be of order 10- 19 J whereas uniform reversal requires energies of
(TrR 2 hHswJs)~6 x
(rrR X 10- 18 - 2 X 10- 17 J. These facts indicate that not only non-
uniform modes (as discussed in Section 8.4.3) but also thermal activation are
involved in the magnetization reversal. A similar behavior has recently been
found in small, pyramidal Ni particles (Ross et al. , 2000).
324 S. Wirth and S. von Monlnar

0.3
Particle
diameter

14 nm
0.1

a 100 200 300


T(K)
Figure 8. 16 Temperature dependence of the mean switching field, H sw ' for different
particle diameters. Low temperature values (T <
(T:< 100 K) are obtained by Hall
magnetometry, whereas room temperature values are taken from MFM. The lines are
obtained by fitting experimental values using Eq. (8.12).
(8. 12).

In order to consider thermal fluctuations influencing the magnetization


reversal process a phenomenological model is introduced. This model does not
rely on the assumption of uniform magnetization orientation for calculating the
energy barrier. Rather, if starts from the concept of a critical volume Vcr
within which a nucleus of reversed magnetization is formed by thermal
fluctuation. Consequently, Vcr has to be separated from the remaining particle
magnetization by a domain wall (Fig. 8.17). Nucleus and wall are successfully
formed if thermal energy and differential gain in magnetostatic energy (inside
Vcr) overcome the wall energy (Givord
CGivord et al., 1988a). We assume that Vcr
covers the complete cross section of the small particles and may vary
somewhat within the array

s Tln( t/ To)
kksTln(t/TO) = cry Tr 1_
"IT d 2
cry _-4- -- HJ
v"fma,
HJ s vCr"'dVcr.
dV cr ' (8.11)
Vcr,min

8
8,j
Vcr \

I
8,
~
k
8H ,
8f1
I
I
I
Figure 8. 17 Schematic illustration of the model for magnetization reversal in elongated
particles. A nucleus of reversed magnetization (volume Vcr) is formed by an applied field
as well as thermal activation. Once the wall is successfully formed (shaded area) it runs
through the complete particle.
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 325

Here, y refers to the wall energy density and a ex accounts for the number of
walls formed as well as for shape inhomogeneities. The thermal energy can be
estimated within the critical volume approach (Chantrell et ai., aI., 1986;
Wegrowe et ai.,aI., 1999) due to the sharp peak in its distribution. The mean
value Vcr
Vcr of the critical volume is thus related to the measured switching field

(8.12)
(8. 12)

A temperature and field independent II


Vcrcr can be assumed due to the weakly

temperature dependent material properties (II (Vcr ~) and is


cr is related to /7f1T{';)

verified by independent experimental findings (li (Li et ai.,


aI., 1997a; Wacquant
aI., 1999). It should be noted that this model can be compared favorably to
et ai.,
analytic calculations (Braun, 1994) and numerical simulations (Boerner and
Bertram, 1997; Brown et al. , 2000; Hinzke and Nowak, 2000). Specifically,
the latter simulation clearly shows the formation of the end caps and the
corresponding domain wall for the smallest particle diameter considered.
Further independent checks of this model will be presented in Sections 8. 5. 3
and 8.5.4.
Although Eq. (8.12)
(8. 12) is found from a simple, plausible model it can also be
derived by an Arrhenius law Eq. (8.9) with E B consisting of magnetostatic and
wall energy.
Eq. (8. 12) predicts an almost Iinearlinear decrease of H sw sw with increasing
temperature. Such a dependence is indeed found experimentally with some
deviations for the biggest (19 nm) particles. The lines in Fig. 8.16 are
obtained by fitting the experimental Hsw-values to Eq. (8. 12). Here, the
temperature dependence of J ss (T) ( n is taken into account. Values of Vcr Vcr can be
estimated from these fits and are listed in Table 8. 1. Note that II Vcrcr is mainly
determined by the temperature dependence of H sw, sw' i. e. the second term of
Eq.(8.12).
The first term in Eq. (8. 12) is more difficult to evaluate. Since the exact
configuration of the domain wall is not known, y can only be approximately
estimated. However, for particles as small as ours, y is expected to become
relatively insensitive to the wall configuration and the particle dimension
(Kneller, 1962; Braun, 1994). Using the bulk value for iron, y = 1. 3 x
2
10 - 33 Jm-
1O- Jm - 2,, we find a
ex ranging from O. 7 - 1. 1 for the different particle arrays.
0.7-1.1
This result is in agreement with the assumption of one wall formed in the
nucleation process and running through the particle. Deviations from the value
a = 1 may be due to particle inhomogeneities or uncertainty in y.
ex
The assumption of a nucleus covering the particle's cross-section yields a
critical length I crcr of the nucleus, IIcr
Vcr = ; d 2 I cr'
cr' Hence, the temperature

dependence of H sw ( T) is expected to be influenced by the particle cross-


section (second term in Eq. (8.12.
(8. 12. Such behavior is experimentally found
for particles with d ~ II cr! d 2 in
::::;;; 17 nm, as can be seen from the constant ratio Veri
326 S. Wirth and S. von Monlnar
Monlncr

Table 8. 1. As expected, the smaller the particle, the stronger the impact of
thermal activation. In fact, our smallest particles (d ~ 9 nm) are
superparamagnetic at room temperature, i. e. , the energy thermally provided
is sufficient to cause magnetization reversal. The 10 nm-particles are very
close to superparamagnetic behavior at room temperature. These findings
(Table 8. 1) are in excellent agreement with calculations by Brown (1963)
who gave the upper volume limit of a superparamagnetic iron sphere to be
270 nm 3 . Our results are also consistent with those on Ni-particles (Ross et
al., 2000). For the 9 nm-particles, the longitudinal relaxation time is
estimated (Coffey et al. , 1994) to be in the order of 10 10-- 33 s, which could be
somewhat reduced by the complex path in configuration space along which the
magnetization reverses (as discussed in Section 8.5. 8. 5. 1).
1) .
The above results support an interpretation that II Verer is the minimum volume
for a nucleus to initiate magnetization reversal in the particles. In addition, the
constant value of V lI erer /d 2 implies a constant ler of about 3.3 nm. The nuclei
extend to about the same length into the particles independent of d (for d :s;;; ~
17 nm) and can most likely be related to the end caps found to initiate the
magnetization reversal in numerical simulations (Brown et al. , 2000; Hinzke
and Nowak, 2000). The value of I er ~ 3.3 3. 3 nm appears plausible since it is just
above lI = ex =
i\ ex 3. 1 nm (Section 8. 4. 3): no inhomogeneous magnetization
configuration is expected on a length scale below II i\ ex .
Another experimental finding can easily be explained within the spirit of
our model. Once the nucleus and the corresponding domain wall are
successfully formed (i. e. the nucleus acquired sufficient volume and does not
collapse), it runs through the complete particle. Accordingly we do not find
any experimental hint of the influence of particle height h on magnetization
reversal within the range 50 nm :s;;; ~ h :s;;;
~ 250 nm investigated by us
experimentally. Smaller particles that do not support a domain wall are
expected to reverse their magnetization uniformly (Bonet et al. , 1999).
The particles with 19 nm in diameter show some deviations from the
above model (see Fig.8.16 Fig. 8. 16 and Table 8. 1). Using constant times t for all
H sw (T) -measurements, the thermal energy at low temperature may not have
been sufficient to provide nuclei with sufficient probability to induce the
nucleation/ propagation process. This may have caused the increased H sw (T)
for these particles as well as the slightly sl ightly enlarged H sw sw for the 17 nm-particles
at T :s;;; ~ 30 K. Alternatively, the deviations of H sw (T) from the expected
straight Iine line for the 19 nm-particles may be indicative of a different reversal
mode from that assumed in our model. In fact, numerical simulations (Hinzke
and Nowak, 2000) for larger particle diameter show a combination of
nucleation at the particle ends and curling at the particle's central region. The
increased number for II Vercr// d 2 in the case of the 19 nm-particles may also
indicate that the critical volume for particles of larger diameter scales to a
higher order than d 2 (d 3 is reported by Li et al. (1997a (1997 a Our assumption that
IIif ercr covers the particles' cross-section certainly breaks down for larger
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 327

diameters.
The shift of H sw
sw ( n toward lower values for the 9 nm-particles is caused
(T)
by more pronounced morphological inhomogeneities in these smallest
particles. These inhomogeneities are reflected in the smallest value, aex = 0.7,
0 . 7,
of all the investigated arrays.
For magnetic data storage, the magnetization has to be stable for
10 years with a retained fraction of O. 95 (Weller and Moser, 1999). Using
(8. 10) we can now calculate the maximum coercivity, H~, under these
Eq. (8.10)
storage conditions (Table 8. 1). For the 9 nm-particles a retained fraction of
0.95 is only kept for tenths of seconds. Moreover, the minimum (write) fields
for reversal in a field pulse of 1 ms, H~, are calculated. H~ roughly follows
sw ( T = 0) due to the low impact of thermal activation at short times.
Hsw(T
H

8.5.3 Angular Dependence of Magnetization Reversal


In order to further establish the credibility of our model we extend it to include
the angular dependence of magnetization reversal. This generalized case has
already been depicted in Fig. 8. 17. For non-colinear H app and EMD, the
particle magnetization is rotated toward the field direction as shown in Section
8. 4. 1. The corresponding angles of the magnetization before and after
e e
reversal are denoted 8 ji and 8 j ' respectively. They can be calculated from
e
Eq. (8. 6) whereby 8 ji ~ "ITTT /2 and 8 j ~ "IT e
TT /2. Note that these angles also
describe the magnetization orientation within II IIcrcr and in the unmodified volume
of the particle while nucleation takes place (Givord et aI., al. , 1988b). The total
angle of the wall separating these volumes is then reduced to 8 ij -- 8 j ~ "IT. e
TT. e
Assuming a linear dependence of wall energy on total wall angle one finds in
analogy to Eq. (8. 12)
1 _ (8 j -8 j )exyd 2 25k s T
-2 Hsw[cos(8 H - 8j ) - cos(8 H - 8j)] - 4J V - -V J .
IIcrJ .
s cr cr s5

(8. 13)

The angular dependence of switching, H sw (e (8 H ) , as calculated from


Eq. (8. 13) is compared to experimental values in Fig. 8. 18. It should be
emphasized that, with II exy already determined from the
IIcrcr and the factor ay
temperature dependence of H sw ', there is no free parameter involved in our
model Eq. (8.13). Hence, the agreement between data measured at 100 K for
the 14 nm- and 17 nm-particles and the results from Eq. (8. 13) strongly
supports our model. One might argue that magnetic defects inside the particles
(Richter, 1989) or interparticle interactions (Wirth, 1995) may lead to
qualitatively similar behavior as the experimentally measured H sw (8 H ) . These
HSW(e
effects, however, are very unl ikely from quantitative considerations (only
shape anisotropy is involved and interparticle interaction fields are estimated
not to exceed the order of 1 mT).
328 S. Wirth and S. von Manlncr
Monlnar

0.6 /
OIOK / !
10
0100 K Curliny 0/
FOA
E0.4 14nm ,96
- 96
-" // ..()/
-0-'<>'
~
~ - <xr-
- -=:....-
;; O<r --<>-
--0- -<:Y
~
...... _
- -<:r- ~
---0--- -- :..r-
~::l. 0.2 q-_-"-_~_--o--
<;>-_...,i;l.._~_~-- Thernal
activation

0.0 ~_.,-':-_"""""_---,'=--_:'=--_-=,"="_---!
'--_'"'=-_--::-'-:-_---,L,..-----,~-~-___!
o 15 30 45 60 75 90
8BH (")
C)
(a)

0.6
o 17nm
17 nm at 100
lOOKK
t:. 19 nm at 75 K
E OA
E 0.4
:<
j~" 0.2
0.0~--c'-=--"""""-----7=---:'=----==-----='
0.0 ',--c'-=----::-'-:------,',,..-----,!-=---='=--___='
o 15 30 45 60 75 90
8H C)
B (")

(b)

Figure 8. 18 Angular dependence of the switching field for different particle diameters and
temperatures. (a) Experimental data for the 14 14nm-particles:
nm-particles: the low- T data are
low-T
compared to the curling (dashed-dot line) and Stoner-Wohlfarth model (dotted line)
whereas data at 100 K can favorably be compared to the thermal activation model (solid
line). The latter applies also for the 17 nm- particles (circles and solid line) in (b). The
thermal activation model does not properly describe the magnetization reversal in the
19 nm-particles (triangles
<triangles and dashed line).

For the 19 nm-particles, there is some systematic deviation between


model and experimental data at 8 H ? "?- 45. These deviations might support the
assumption of a different reversal mode in these biggest particles as already
discussed for HswCT) in the previous section.
The values H sw (8 C8 H)
H ) approach H a for 8 H n/2.
H --.. TI /2. Such behavior is
expected for small, single-domain particles and is well reproduced by
Eq. (8.13).
(8. 13). This also causes the crossing of the calculated H sw (8 C8 H ) in
Fig. 8. 18b since the 19 nm-particles (dashed line) exhibit a slightly larger
shape anisotropy than the 17 nm-particles Csolid (sol id line). Note that the
experimentally determined switching field distribution of the 17 nm-particles
CSection 8.4.1
(Section 8.4. 1 and Fig.8.1D
Fig. 8. 11) has been taken into account in calculating mean
Hsw (8
values of H C8 H ) for this specific array (Eq.
CEq. (8.
C8. 13) and solid line in Fig. 8. 18b).

8.5.4 Magnetic Viscosity


As another independent check on our magnetization reversal model, magnetic
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 329

viscosity experiments are performed, i. e. , the time dependence of the total


array magnetization is investigated. It should be emphasized that these
measurements are only possible in the optimized Hall configuration (Section
8.3.3) and are, nonetheless still at the limits of sensitivity due to the
miniscule changes in total magnetization.
The principle of magnetic viscosity measurements is depicted in
Fig. 8. 19. After initial saturation of the sample at positive fields, H = 0, the e
field is reversed close to H sw . This reduces the energy barrier for
magnetization reversal significantly and those particles with a smaller energy
barrier reverse their magnetization (demonstrated by the drop of
magnetization starting at H app.......
app""'" - O. 13 T in Fig. 8. 19). The field is then kept

constant and the time dependence of total magnetization is measured. After


that, the field is swept back to positive saturation to detect any drift in the
experimental setup which would manifest itself in a non-closed loop.

0.4

.
~
55' .2
00.2 ; {/:t
I
<1 d=17nm
d=17 nm
. :t;:~.1 ~~.l
~ ;:,:"".1 ... ..:'\
,::;../....,/' T=20 K
"t'...
....'f..
~.
~

'~""'" ,w:::':
0
'0
,W:-:.:
.
.......
. ,.
""H=-O . 17T
rv
0.0

--0.2
--D.2 -0.1
--D. I 0.0 0.1 0.2 0.3 0.4
J.loHapp(T)
lloHapp(T)

Figure 8. 19 Schematic illustration and check of stability of experimental parameters for


the magnetic viscosity measurements (array of 17 nm particles). The time dependent
magnetization measurements are conducted at reversing fields of - O. 17 T. This field
value is marked (arrow) for the sweep back to positive fields.

The dependence of the total array magnetization on time can, again, be


described by Eq. (8. 11). It predicts a logarithmic dependence as long as the
distribution of energy barriers (i. e., of II cr over which the integration is
performed) is sufficiently broad and smooth (Neel, 1949): J ( t) = = J (0) - S .
tin with S being the magnetic viscosity. Such behavior is indeed found
In( tit')
for the 17 nm-particles and intermediate time ranges. An example measured at
= 140 mT (close to the mean switching field) is shown in
75 K and 1-10 H app =
Fig. 8.20, main panel. About 14 particles reverse their magnetization within
the time range shown although the reversal of individual particles cannot be
resolved. Using different conditions (inset in Fig. 8. 20, see also (Lederman et
al. , 1994 the onset and trail-off of magnetization reversal is observed. This
behavior is determined by the integral in Eq. (8. 11). The line (inset in
Fig. 8.20) is calculated from an exponential law for the switching
probabilities.
330 S. Wirth and S. von Monlnar

0.1
0.0
Diameter
1711111
17nm -0.]
0.0 --D. I
T=75 K
-{) .~- ' - -_
-0.2 llo H_
JloH=-185mT
=-185_
mT_.c...,..._
>
:; -0.1
:::1.
~
10
~

100 ]I 000

"'"

-{U
T=75 K. J1 o. H=-140 111T
K, lloH=-140 mT
-o.3':-:::--~~-~~~--:-:::'::-::---.".-,J
200 1 000 2 000
Time(s)
Time (s)

Figure 8. 20 Time dependence of the total magnetization of the 17 nm-particle array in


consequence of thermal activation. For intermediate time and field range (main panel) a
logarithmic time dependence of the magnetization is observed. Different conditions reveal
also the onset and trail-off of the individual particle magnetization reversal (inset); the line
is calculated from an exponential law of switching probabilities.

At constant H app
app the magnetization reversal is caused by mere thermal

activation. This reversal, which acts in addition to the field-driven one, is


8. 19. For upward sweeping field, H app
depicted in Fig. 8.19. app is indicated by an

arrow. A somewhat higher field is needed to regain saturation in order to


compensate for the thermally activated reversal contribution. Since it certainly
may be assumed that the same energy barriers have to be overcome for field-
driven and thermally activated reversal, the comparison of these two types of
activation yields a so-called activation volume

v = k B T JJ 0 X irr (8 14)
(8.
a Js S

which gives the typical volume involved in a single activation process


X,rr (i
(Wohlfarth, 1984). The irreversible susceptibility Xirr (i. e. the irreversible
change in magnetization) is taken from magnetization curves H = O. For the e
conditions of Fig. 8 Va~250
8. 20 we find V 3
a =250 nm . Similar results were obtained in

the temperature range 10 - 60 K. Considering the different concepts involved,


Va and Vcr are found to be in reasonable agreement. The magnetic viscosity
experiments confirm the obtained values of the critical volume
volume.
Our results for Va are somewhat smaller if compared to values from
literature (li
(Li et al., 1997a; Stinnett and Doyle, 1998; Wacquant et al.,
1999). In contrast to those, however, we investigate an extremely small total
magnetic volume of identical particles. Thus interparticle interactions and
particle misalignments could be excluded to a large extend. Our measurements
limit of sensitivity, on the other hand, may result in dim
at the Iimit diminished
inished
accuracy.
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 331

8. 6 Conclusions

We fabricate arrays of iron particles by 8TM-assisted CVD. This


technique is able to produce well shaped nanoparticles of controllable
dimensions at predefined positions. The latter is a prerequisite for optimized
Hall measurements that allow a detailed study of magnetization processes of
such particles. Our fabrication and measurement techniques are briefly
compared to other methods suitable for nanoparticles.
The shape anisotropy of the particles can be determined from an analysis
of the reversible processes which take place uniformly. In contrast, the
irreversible processes are of non-uniform nature. The latter result has severe
consequences for small particle magnetic storage media for which
homogeneous reversal is commonly assumed.
The decrease in switching field with increasing temperature due to thermal
activation is explained within a phenomenological model which is confirmed by
angular dependent and viscosity measurements. In our elongated particles of
9 - 17 nm diameter, magnetization reversal is initiated by nucleation. If a
nucleus of reversed magnetization is successfully formed the corresponding
wall immediately runs through the whole particle. Hence, the critical volume of
the nucleus scales with the particle cross-section and is much smaller than the
particle volume. This model breaks down for larger particle diameter. Our
smallest particles are superparamagnetic at room temperature.
Our model certainly can be adapted to other particle geometries. It will
break down if all three dimensions of the particles are reduced to or below the
wall width. In this case, homogeneous magnetization reversal is expected.

References
Aharoni, A. Phys. 8tat. 801. 16: 3 (1966)
Aharoni, A. Introduction to the Theory of Ferromagnetism. Oxford University
Press, Oxford, (1996)
(1997)
Arias, R. and H. N. Bertram. J. Magn. Magn. Mater. 171: 209 (1997>
Babcock, K. L. , V. B. Elings,
EI ings, M. Dugas and 8. Loper. IEEE Trans. Magn.
30: 4503 (1994)
Babcock, K. L., V. B. Elings, J. 8hi, D. D. Awschalom and M. Dugas.
Appl. Phys. Lett. 69: 705 (1996)
8. J. , Advances in Physics 48: 449 (1999)
Bending, 8.J.,
Bertram, H.N. Peng. IEEE Trans. Magn. 34: 1543 (1998)
H. N. and Q. Pengo
Billas, I. M. L. , J. A. Becker, A. Chatelain and W. A. de Heer. Phys. Rev.
Lett. 71: 4067 (1993)
332 S. Wirth and S. von Monlnar

Binnig, G., H. Rohrer, C. Gerber and E. Weibel. Phys. Rev. Lett. 49: 57
(1982)
Boerner, E. D. and H. N. Bertram. IEEE Trans. Magn. 33: 3052 (1997)
Bonet, E., W. Wernsdorfer, B. Barbara, A. Benoit, D. Mailly and A.
Thiaville A. Phys. Rev. Lett. 83: 4188 (1999)
H. -B. Phys. Rev. B
Braun, H.-B. 8 50: 16,501 (1994)
Brown, G., M. A. Novotny and P. A. Rikvold. J. Appl. Phys. 87: 4792
(2000)
Brown, W.F, Jr. Phys. Rev. 130: 1677 (1963)
Cadieu, F. J. , C. Caldwell, J. Griffin and S. von Molnar. J. Appl. Phys. 81:
5082 (1997)
Chambliss, D. D. , R.J.
D.O., R. J. Wilson and S. Chiang. J. Vac Sci. Technol. B9: 89:
933 (1991)
Chantrell, R. W. , M. Fearnon and E. P. Wohlfarth. Phys. Status Solidi A 97:97 :
213 (1986)
Chantrell, R. W. and J. Hannay. J. Magn. Soc. Jpn. 21: 283 (1997)
Chapman, J. N. and M. R. Scheinfein. J. Magn. Magn. Mater. 200: 729
(1999)
( 1999)
Chappert, C. , H. Bernas, J. Ferre, V. Kottler, J. -P. Jamet, Y. Chen, E.
Cambril, T. Devolder, F. Rousseaux, V. Mathet and H. Launois. Science
280: 1919 (1998)
Choi B.C.,
B. C. , M. Belov, W.K.
W. K. Hiebert, G.E.
G. E. Ballentine and M.R.
M. R. Freeman.
Phys. Rev. Lett. 86: 728 (2001)
Chou, S.Y., M.S. Wei, P.R. KraussandP.B. Fischer. J. Appl. Phys. 76:
6673 (1994a)
Chou, S. Y., M. S. Wei, P. R. Krauss and P. B. Fischer. J. Vac. Sci. Sci
Technol. B 8 12: 3695 (1994b)
Chou, S. Y. , P. R. Krauss and P. J. Renstrom. Science 272: 85 (1996)
Christoph V., S. Wirth and S. von Molnar. J. Appl. Phys. 89:7472 (2001)
Coffey, W. T., P. J. Cregg, D. S. F. Crothers, J. T. Waldron and A. A W.
Wickstead. J. Magn. Magn. Mater. 131: L301 (1994)
Coffey, W. T.,T. , D. S. F. Crothers, J. L. Dormann, L. J. Geoghegan, Y. P.
Kalmykov, J. T. Waldron and A. W. Wickstead. Phys. Rev. B 8 52: 15,951
((1995)
1995)
Cowburn, R. P. , A. O. Adeyeye and J. A. C. Bland. Appl. Phys. Lett. 70:
2309 (1997)
Dan Dahlberg E. and R. Proksch. J. Magn. Magn. Mater. 200: 720 (1999)
de Lozanne, A. Jpn. J. Appl. Phy. 33, pt. 1: 7090 (1994)
de Lozanne, A. Supercond. Sci. & & Technol. 12: R43 (1999)
Dunin-Borkowski, R. E. , M. R. McCartney, B. Kardynal, D. J. Smith and M.
R. Scheinfein. Appl. Phys. Lett. 75: 2641 (1999)
Duvail, J. L., S. Dubois, L. Piraux, A. Vaures, A. Fert, Fer!, D. Adam, M.
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 333

Champagne, F. Rosseaux and O. D. Dacanini.


Oacanini. J. Appl. Phys. 84: 6359
((1998
1998))
Ehrichs, E. E.,E. , R. M. Silver and A. L. de Lozanne. J. Vac. Sci. Technol.
A 6: 540 (1988)
0' Grady. J. App. Phys. 84: 5114
EI-Hilo, M., R. W. Chantrell, and K. O'Grady.
(1998)
Fernandez, A, P. J. Bedrossian, S. L. Baker, S. P. Vernon and D. O. R. Kania.
IEEE Trans. Magn. 32: 4472 (1996)
Field, M. unpublished data (1998)
Folks, L., M. E. Best, P. M. Rice, B. D. O. Terris, D. O. Weller and J. N.
Chapman. Appl. Phys. Lett. 76: 909 (2000)
Fredkin, O.R.D.R. and
andT.R.
T.R. Koehler. J. Appl. Phys. 67: 5544 (1990)
Frei, E.H.,
E. H. , S. Shtrikman and O.D. Treves. Phys. Rev. 106: 446 (1957)
Dubonos, J. G. S. Lok, I. V. Grigorieva, J. C. Maan, L.
Geim, A. K. , S. V. Oubonos,
Theil Hansen and P.E.P. E. Lindelof. Appl. Phys. Lett. 71: 2379 (1997a)
Geim, A. A.K., I. V. Grigorieva, S.V.
K. , I.V. S. V. Dubonos,
Oubonos, J.G.S.
J. G. S. Lok, J.C.
J. C. Maan, A.
E. Filippov and F.M.F. M. Peeters. Nature 390: 259 (1997b)
Gider, S., J. Shi, D.D. D.O. Awschalom, P.F. Hopkins, K.L. Campman, A.C.
Gossard, A.D.A. D. Kent and S. von Molnar. Appl. Phys. Lett. 69: 3269
(1996)
Givord, D., P. Tenaud and T. Viadieu. IEEE Trans. Magn. 24: 1921
( 1988a)
Givord, D., P. Tenaud and T. Viadieu. J. Magn. Magn. Mater. 72: 247
(1988b)
Goddenhenrich, T. , H. Lemke, M. Muck, U. Hartmann and C. Heiden. C.
G6ddenhenrich,
Appl. Phys. Lett. 57: 2612 (1990)
Gomez, R. O. D. , T. V. Luu, A. O. Pak, K. J. Kirk and J. N. Chapman. J.
Appl. Phys. 85: 6163 (1999)
Grutter, P., H. J. Manim and D.
GrOtter, O. Rugar. In: Wiesendanger and H. -J.
GOntherodt eds. Scanning Tunneling Microscopy II, Springer, p. 151
Guntherodt
(1995)
Haast, M. A. M. , I. R. Heskamp, L. Abelmann, J. C. Lodder and T. J. A.
Popma. J. Magn. Magn. Mater. 193: 511 (1999)
Haginoya, C, S. Heike, M. Ishibashi, K. Nakamura, K. Koike, T.
Yoshimura, J. Yamamoto and Y. Hirayama. J. Appl. Phys. 85: 8327
(1999)
Hallen, H.O.,
H. D. , R. Seshadri, A.
A.M.
M. Chang, R.E.
R. E. Miller, L.N.
L. N. Pfeiffer, K.W.
K. W.
West, C. A. Murray and H. F. Hess. Phys. Rev. Lett. 71: 3007 (1993)
Hiebert, W. K. , A. Stankiewicz and M. R. Freeman. Phys. Rev. Lett. 79:
1134 (1997)
Hinzke, O. D. and U. Nowak. J. Magn. Magn. Mater. 221: 365 (2000)
Hirayama, T., Q. Ru, T. Tanji and A. Tonomura. Appl. Phys. Lett. 63: 418
334 S. Wirth and S. von Monlnar

(1993).
Hubert, A. and R. Schafer. Magnetic Domains. Springer, Berl in,
Heidelberg, New York (1998)
Ichihara, K., A. Kikitsu, K. Yusu, F. Nakamura and H. Ogiwara. IEEE
Trans. Magn. 34: 1603 (1998)
Ju, G. , A. V. Nurmikko, R. F. C. Farrow, R. F. Marks, M. J. Carey and B.
A. Gurney. Phys. Rev. Lett. 82: 3705 (1999)
Kent, A. D., T. M. Shaw, S. von Molnar and D. D. Awschalom. Science
262: 1249 (1993)
A.D.,
Kent, A. D. , S. von Molnar, S. Gider and D.D.
D. D. Awschalom. J. Appl. Phys.
76: 6656 (1994)
Kirk, K.J.,
K. J. , J.
J.N.
N. Chapman
ChapmanandC.D.W.
and C. D. W. Wilkinson. J. Appl. Phys. 85: 5237
(1999)
Kirsch, S., A. Pollmann , M. Thielen, H. Weinforth, A. Carl, E. F.
Wassermann. J. Appl. Phys. 81: 5474 (1997)
Kirtley, J.R.,
J. R. , M.B.
M. B. Ketchen, K.G.
K. G. Stawiasz, J.
J.Z.
Z. Sun, W.
W.J.
J. Gallagher,
S.H.
S. H. Blanton and S.J.
S. J. Wind. Appl. Phys. Lett. 66: 1138 (1995)
Kirtley, J. R., C. C. Tsuei, J. Z. Sun, C. C. Chi, L. S. Yu-Jahnes, A.
Gupta, M. Rupp and M. B. Ketchen. Nature 387: 481 (1997)
Kleiber, M., F. Kummerlen, M. Lohndorf, A. Wadas, D. Weiss and R.
Wiesendanger. Phys. Rev. B 58: 5563 (1998)
Kneller, E. Ferromagnetism. Springer, Berlin, pp.292-300 (1962)
Kong,L. , L. Zhuang and S. Y. Chou. IEEE Trans. Magn. 33: 3019 (1997)
Kreuzer, S. , K. Prugl, G. Bayreuther and D. Weiss. Thin Solid Films 318:
219 (1998)
Kurkijarvi, J. Phys. Rev. B 6: 832 (1972)
Lambeth, D. N. , E. M. T. Velu, G. H. Bellesis, L. L. Lee and D. E. Laughlin.
J. Appl. Phys. 79: 4496 (1996)
Lederman, M. , S. Schultz and M. Ozaki. Phys. Rev. Lett. 73: 1986 (1994)
Leinenbach, P., U. Memmert, J. Schelten and U. Hartmann. Appl. Surf.
Sci. 145: 492 (1999)
Li, F., R. M. Metzger and W. D. Doyle. IEEE Trans. Magn. 33: 3715
(1997a)
Li, N. and B. M. Lairson. IEEE Trans. Magn. 35: 1077 (1999)
Li, S.P., W.S. Lew, Y.B. Xu, A. Hirohata, A. Samad, F. BakerandJ.A.
C. Bland. Appl. Phys. Lett. 76: 748 (2000)
Li, X.-Q., F. M. Peeters and A. K. Geim. J. Phys.: Condens. Matter 9:
8065 (1997b)
Lohau, J. , S. Kirsch, A. Carl A. G. Dumpich and E. F. Wassermann. J.
Appl. Phys. 86: 3410 (1999)
Lottis, D., F. Petroff, A. Fert and M. Konczykowski. J. Magn. Magn.
104 -7: 1811 (1992)
Mater. 104-7:
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 335

Lyberatos, A.,
A. , R. W. Chantrell and A. Hoare. IEEE Trans. Magn. 26: 222
(1990)
M. R. Scheinfein and J. M. Cowley. J. Appl. Phys. 75: 7418
Mankos, M., M.R.
(1994 )
Martin, J. I., Y. Jaccard, A. Hoffmann, J. Nogues, J. M. George, J. L.
Vicent, I. K. Schuller. J. Appl. Phys. 84: 411 (1998)
Matteucci, G., M. Muccini and U. Hartmann. Appl. Phys. Lett. 62: 1839
(1993)
McCord, M. A. and D. D. Awschalom. Appl. Phys. Lett. 57: 2153 (1990)
Meier, G., D. Grundler, K. -B. Broocks, C. Heyn and D. Heitmann. J.
Magn. Magn. Mater. 210: 138 (2000)
Meier, J. , B. Doudin and J. P. Ansermet. J. Appl. Phys. 79: 6010 (1996)
Monzon, F. G. , D. S. Patterson and M. L. Roukes. J. Magn. Magn. Mater.
195: 19 (1999)
Nakayama, M., F. Wakaya, J. Yanagisawa and K. Gamo. J Vac. Sci.
Technol. B 16: 2511 (1998)
Neel, L. Ann. Geophys. 5: 99 (1949)
New, R. M. H., R. F. W. Pease and R. L. White. J. Vac. Sci. Technol.
B 12: 3196 (1994)
New, R. M. H., R. F. W. Pease and R. L. White. J. Vac. Sci. Technol.
B 13: 1089 (1995)
Nowak, U. , R. W. Chantrell and E. C. Kennedy. Phys. Rev. Lett. 84: 163
(2000)
Ohresser, P., G. Ghiringhelli, O. Tjernberg and N.B. Brookes. Phys. Rev.
B 62: 5803 (2000)
A.,, S.J.
Oral, A. S. J. Bending and M. Henini. Appl. Phys. Lett. 69: 1324 (1996)
Park, Y. D. , J. A. Caballero, A. Cabbibo, J. R. Childress, H. D. Hudspeth,
T. J. Schultz and F. Sharifi. J. Appl. Phys. 81: 4717 (1997)
T.J.
and M. Pepper. J. Appl. Phys. 87: 3171 (2000)
Patel, N.K. andM.
Peeters, F. M. and J. De Boeck. In: H. S. Nalwa ed. Handbook of
nanostructured materials and nanotechnology. Academic
Academ ic Press, New York,
Vol. 3 p. 345 (1999)
Porthun, S. , L. Abelmann and C. Lodder. J. Magn. Magn. Mater. 182: 238
(1998)
Proksch, R., E. Runge, P. K. Hansma, S. Foss and B. Walsh. J. Appl.
Phys. 78: 3303 (1995)
Proksch, R. Curro Opin. Solid State Mat. Sci. 4: 231 (1999)
Oiu,
Qiu, Z.O.
Z. Q. and S.D.
S. D. Bader. J. Magn. Magn. Mater. 200: 664 (1999)
Rave, W. , K. Ramstock and A. Hubert. J. Magn. Magn. Mater. 183: 329
(1998)
Richter, H. J. J. Appl. Phys. 65: 3597 (1989)
Ross, C.A., R. Chantrell, M. Hwang, M. Farhoud, T. T.A.
A. Savas, Y. Hao,
336 S. Wirth and S. von Monlnar

H. I. Smith, F. M. Ross, M. Redjdal and F. B. Humphrey. Phys. Rev.


H.1.
B 62: 14,252 (2000)
Rousseaux, F., D. Decanini, F. Garcenac, E. Cambril, M. F. Ravet, C.
Chappert, N. Bardou, B. Bartenlian and P. Veillet. J. Vac. Sci. Technol.
B 13: 2787 (1995)
Rubel, S. , M. Trochet, E. E. Ehrichs, W. F. Smith and A. L. de Lozanne. J.
Vac. Sci. Technol. B 12: 1894 (1994)
Rugar, D., H. J. Mamin, P. Guethner, S. E. Lambert, J. E. Stern, I.
McFadyen and T. Yogi. J. Appl. Phys. 68: 1169 (1990)
Ruhrig, M.,
ROhrig, M. , B. Khamsehpour, K. J. Kirk, J. N. Chapman, P. Aitchison, S.
McVitie and C. D. W. Wilkinson. IEEE Trans. Magn. 32(5) 32 ( 5) pt. 2: 4452
(1996a)
Ruhrig, M. , S. Porthun, J. C. Lodder, S. McVitie, L. J. Heyderman, A. B.
ROhrig,
Johnston and J. N. Chapman. J. Appl. Phys. 79: 2913 (1996b)
Runge, K., Y. Nozaki, Y. Otani, H. Miyajima, B. Pannetier, T. Matsuda
and A. Tonomura. J. Appl. Phys. 79: 5075 (1996)
Saifullah, M. S. M. , C. B. Boothroyd, C. J. Morgan and C. J. Humphreys.
Electron Microscopy and Analysis. lOP PublPubIishing,
ishing, Bristol, p. 325 (1995)
Savas, T. A. , M. Farhoud, H.1.
H. I. Smith, M. Hwang and C. A. Ross, J. Appl.
Phys. 85: 6160 (1999)
Schabes, M.E.
M. E. J. Magn. Magn. Mater. 95: 249 (1991)
Scheinfein, M. R. , J. Unguris, J. L. Blue, K. J. Coakley, D. T. Pierce, R. J.
Celotta and P. J. Ryan. Phys Rev. B 43: 3395 (1991)
Schrefl, T. J. Magn. Magn. Mater. 207: 66 (1999)
Schwarzacher, W., K. Attenborough, A. Michel, G. Nabiyouni and J. P.
Meier. J. Magn. Magn. Mater. 165: 23 (1997>(1997)
Schweinbock,
Schweinbbck, T. , D. Weiss, M. Lipinski and K. Eberl. J. Appl. Phys. 87:
6496 (2000)
Skomski, R., J.P. LiuandD.J. Sellmyer. Phys. Rev. B60: 7359 (1999)
Skomski, R. , H. Zeng, M. Zheng and D. J. Sellmyer. Phys. Rev. B 62:
3900 (2000)
Smith, H. I., D. J. D. Carter, M. Meinhold, E. E. Moon, M. H. Lim, J.
Ferrera, M. Walsh, D. Gil and R. Menon. Microelectronic Engineering 53:
77 (2000)
Smyth, J. F., S. Schultz, D. R. Fredkin, D. P. Kern, S. A. Rishton, H.
Schmid, M. Cali and T.R. Koehler. J. Appl. Phys. 69: 5262 (1991)
Stavroyiannis, S., I. Panagiotopoulos, D. Niarchos, J. A. Christodoulides,
Y. Zhang and G. C. Hadjipanayis.
Hadj ipanayis. J. Magn. Magn. Mater. 193: 181
( 1999)
Stinnett, S.M. and W.D.
W.O. Doyle. IEEE Trans. Magn. 34: 1681 (1998)
Stoner, E. C. and E. P. Wohlfarth. Philos.
Phi los. Trans. R. Soc. London A 240:
599 (1948)
Fabrication and Magnetic Properties of Nanometer-Scale Particle Arrays 337

Streblechenko, D. G. , M. R. Scheinfein, M. Mankos and K. Babcock. IEEE


Trans. Magn. 32: 4124 (1996)
Suhl, H. and H. N. Bertram. J. Appl. Phys. 82: 6128 (1997)
Sun, S. , C. B. Murray, D. Weller, L. Folks and A. Moser. Science 287:
1989 (2000)
Terris, B. D., L. Folks, D. Weller, J. E. E. Baglin, A. J. Kellock, H.
Rothuizen and P. Vettiger. Appl. Phys. Lett. 75: 403 (1999)
Thiaville, A., L. Belliard, D. Majer, E. Zeldov and J. Miltat. J. Appl.
Phys. 82: 3182 (1997)
Thien Binh, V. , S. T. Purcell, V. Semet and F. Feschet. Appl. Phys. Lett.
72: 975 (1998).
Thomas, L. , F. Lionti, R. Ballou, D. Gatteschi, R. Sessoli and B. Barbara.
Nature 383: 145 (1996)
Tomlinson, S.L. and E.W. Hill. J. Magn. Magn. Mater. 161: 385 (1996)
van Schendel P. J. A. , H. J. Hug, B. Stiefel S. Martin and H. -J. Guntherodt.
J. Appl. Phys. 88: 435 (2000)
Victora, R.H. Phys. Rev. Lett. 63: 457 (1989)
Wacquant, F. , S. Denolly, A. Giguere, J. -P. Nozieres, D. Givord and V.
Mazauric. IEEE Trans. Magn. 35: 3484 (1999)
Wegrowe, J. -E. , O. Fruchart, J. -P. Nozieres, D. Givord, F. Rousseaux,
D. Decanini and J. Ph. Ansermet. J. Appl. Phys. 86: 1028 (1999)
Weller, D. and A. Moser. IEEE Trans. Magn. 35: 4423 (1999)
Wernsdorfer, W., K. Hasselbach, D. Mailly, B. Barbara, A. Benoit, L.
Thomas and G. Suran. J. Magn. Magn. Mater. 145: 33 (1995a)
Wernsdorfer, W. , K. Hasselbach, A. Benoit, G. Cernicchiaro, D. Mailly,
B. Barbara and L. Thomas. J. Magn. Magn. Mater. 151: 38 (1995b)
Wernsdorfer, W., B. Doudin, D. Mailly, K. Hasselbach, A. Benoit, J.
Meier, J.-Ph.
J. -Ph. Ansermet and B. Barbara. Phys. Rev. Lett. 77: 1873
((1996)
1996)
Wernsdorfer, W. , E. B. Orozco, K. Hasselbach, A. Benoit, B. Barbara, N.
Demoncy, A. Loiseau, H. Pascard and D. Mailly. Phys. Rev. Lett. 78:
1791 (1997)
Wirth, S. J. Appl. Phys. 77: 3960 (1995)
Wirth, S. , J. J. Heremans, S. von Molnar, M. Field, K. L. Campman
Campman,, A.
C. Gossard and D. D. Awschalom. IEEE Trans. Magn. 34: 1105 (1998a)
Wirth, S. , M. Field, D. D. Awschalom and S. von Molnar. Phys. Rev. B 57 :
R14,028 (1998b)
Wirth, S.,
S. , In: L. SchultzandK.-H.
Schultz and K. -H. Mulleeds.
MOlle eds. Proc. 10 thlnternatl.
th Internatl. Symp.
on Magn. Anisotropy & Coercivity in RE-TM
RE- TM Alloys. Dresden, Germany,
p. 139 (1998)
Wirth, S. , M. Field, D. D. Awschalom and S. von Molnar. J. Appl. Phys.
85: 5249 (1999)
338 S. Wirth and S. von Monlnar

Wirth, S. and S. von Molnar. Appl. Phys. Lett. 76: 3283 (2000a)
Wirth, S. and S. von Molnar. J. Appl. Phys. 87: 7010 (2000b)
Wirth, S.,
S. , A. Anane and S. von Molnar. Phys. Rev. B 63: 012,402 (2001)
E. P. J. Phys. F: Met. Phys. 14: L
Wohlfarth, E.P. 155 (1984)
L155
Wong, J., A. Scherer, M. Todorovic and S. Schultz. J. Appl. Phys. 85:
5489 (1999)
Ye, P. D., D. Weiss, R. R. Gerhardts, M. Seeger, K. von Klitzing, K.
Eberl and H. Nickel. Phys. Rev. Lett. 74: 3013 (1995)
Yu, M. , Y. Liu, A. Moser, D. Weller and D. J. Sellmyer. Appl. Phys. Lett.
75: 3992 (1999)
Zangari, G. and D. N. Lambeth. IEEE Trans. Magn. 33: 3010 (1997)
Zheng, M., M. Yu, R. Skomski, S. H. Liou, D. J. Sellmyer, Y. Liu, V. N.
Petryakov, Yu. K. Verevkin, N. I. Polushkin and N. N. Salashchenko. J.
Appl. Phys. (2001) to appear
Zueco, E., W. Rave, R. Schafer, A. Hubert and L. Schultz. J. Magn.
Magn. Mater. 190: 42 (1998)

Most of this work has been conducted at the Center for Materials Research and Technology
(Martech), Florida State University. The 2DES material was kindly provided by A. C.
Gossard, UCSB. We thank J. R. Childress (UF) for deposition of the permalloy films.
Portions of the MFM measurements were performed by M. Field (UCSB). The BEM
simulation method was developed by V. Christoph (HTW Dresden, Germany). We
particularly wish to thank D. D. Awschalom CUCSB) , with whom we started these studies in
1991. We are also grateful to A. Anane (FSU) and R. Skomski (UNU (UNL) for very helpful
discussions. This work was supported by NSF Grant No. DMR 95-10518-27553 and AFSOR
Grant No. F49629-96-1-0026. S. Wirth gratefully acknowledges support by the Humboldt
foundation, Germany.
9 Processing and Modeling of Novel Nanocrystalline
Soft Magnetic Materials

Kiyonori Suzuki

9. 1 Introduction

One of the most significant effects of reducing the structural correlation length
of ferromagnetic materials to a nanometer scale is magnetic softening. This
softening is brought about by the reduced effect of the intrinsic
magnetocrystalline anisotropy due to the mutual transmission of the anisotropy
energy among a great number of nanocrystallites. This softening effect
provides us with an alternative approach to the development of novel soft
magnetic materials. Nanocrystalline soft magnetic alloys prepared by
annealing melt-spun amorphous precursors, with which this chapter is
concerned, are one of the latest successful outcomes of such a new approach
to the development of novel soft magnetic materials. In this chapter we will
discuss the history, origin of the softness, processing and properties of the
nanocrystall ine soft magnetic alloys with particular attention paid to:
CDnanostructure-magnetic properties relationships and CZ)
(j)nanostructure-magnetic ~ the principles
underlying material design.
Historically, the discovery of the nanocrystalline soft magnetic alloys
seems to have originated from research aimed at the improvement of soft
magnetic properties in amorphous alloys. Although optimum soft magnetic
properties of melt-spun amorphous alloys were usually obtained after stress-
rei ief anneal ing (Luborsky, 1983), it was widely belbelieved
ieved that the magnetic
softness of any amorphous alloy deteriorates by increasing the annealing
temperature far beyond the crystallization temperature. This common belief
was wrong. In 1988, Yoshizawa et al. (1988) of Hitachi Metals discovered
that the magnetic softness of melt-spun amorphous Fens Si 13 . S5 8 9 Nb 3 Cu, is
Nb3CUl
improved significantly by crystallization. This alloy, commercially known as
(/-Ie) of _105
FINEMET, exhibited an exceptionally high effective permeability (/.Ie)
with a saturation magnetization (M ss)) of 1.25
1. 25 T. This improved magnetic
softness originated from the formation of a unique microstructure composed of
extremely fine grains with a size of about 10 nm. In the following year, Herzer
(1989) analyzed the magnetocrystalline anisotropy energy of nanocrystalline
materials based on the concept of the so-called random anisotropy model
340 Kiyonori Suzuki

originally proposed by Alben et al (1978) for amorphous systems. Herzer


demonstrated that this model is applicable to nanocrystalline systems and
pointed out that the coercivity in nanocrystalline materials scales as the 6 th
exponential power of the grain size.
The discovery of the excellent magnetic core properties in nanocrystalline
Fen
Fe735Si135BgNb3Cul
5Si 13 5B9Nb 3CUI initiated an era of intense research on the development of
new Fe-based nanocrystalline soft magnetic alloys in various alloy systems. As
a result, (Cu, Si )-free-Fe-metal based nanocrystalline alloys, known as
NANOPERM, whose Fe content is much higher than the Fe-metalloid based
alloys, such as Feg Zr7 B2
Fe91t Zr) 2 ,, were developed by Suzuki et al. al (1990). The
4
saturation magnetization of nanocrystalline Fe91Zr)B2 /-Ie of 3 x
Fe9l Zr7 B2 with lJe X 10 was as
high as 1.7 T (Suzuki et al., 1996), the highest value among the
nanocrystalline soft magnetic alloys reported to date, while the permeability of
the Fe735 SitU g Nb3Cui alloy with M s of 1.25 T is still the highest in the
Si 135 B9
nanocrystalline soft magnetic alloy family (Fig. 9. 1). Besides these bulk-form
alloys, thin-film nanocrystalline soft magnetic alloys based on similar principles
were also reported for Fe-EM-C (Hasegawa and Saito, 1990) and Fe-EM-N
(Tanekoetal.,
(Taneko et al. , 1991; Ishiwataetal.,
Ishiwata et al , 1991; Shimizuetal., (EM = IVa
Shimizu et al. , 1991) (EM=IVa
to Via metals) systems. An excellent review on the nanocrystalline Fe-EM-C
thin-films can be found elsewhere (Hasegawa et al. , 1993a).

10"
Nanocrystalline

Fe-Si-B-Nb-Cu
o
A o Fe-Zr-B
Co-based
amorphous
o

Fe-based

FelTits
amorphous

Fe-Si steel

(Empirical limit)

1O=!:---;;-'~----c---~---::--------;:'
o 0.5 I 1.5 2 2.5
Saturation induction (T)

Figure 9.1
9. I Saturation induction and relative permeability at 1 kHz in various soft magnetic
materials.

The unique microstructure in the nanocrystalline soft magnetic alloys has


stimulated intensive investigations of the decomposition behavior in the
amorphous precursors. Advanced experimental techniques including in-situ
time-resolved X-ray diffractometry (Koster et al. , 1991), atom-probe field ion
microscopy (APFIM) (Hono et al., 1992; Zhang et al., 1996) and X-ray
absorption fine structure (XAFS) (Kim et al. , 1993) were employed in these
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 341

investigations. These studies showed that the nanostructural evolution in the


Fe-Si-B-Nb-Cu and Fe-Zr-B systems is due to primary crystallization (Koster
and Herold, 1981) of the precursor amorphous phase, and the resultant
nanoscale structure consists of the primary bee-Fe precipitates and the residual
amorphous matrix. This peculiar microstructure as well as the mechanism
behind the magnetic softness is also attracting growing interest from the
viewpoint of the fundamental magnetism problems in heterogeneous
nanostructural systems.
As reviewed by Gleite (1989), the concept of nanocrystalline materials
can be found as early as in 1981. The basic idea of these nanocrystall ine
materials was to generate a new class of disordered solids by introducing a
high density of grain boundary defects. However, the idea of recent
nanocrystalline soft magnetic alloys is to reduce the effects of the
magnetocrystalline anisotropy by simply reducing the grain size below the
exchange correlation length, different from the principal idea of the previous
nanocrystall ine materials. Hence, some authors in the early stages of research
nanocrystalline
on nanocrystalline soft magnetic alloys preferred to refer to their materials as
ultrafine grain or microcrystalline, rather than nanocrystalline. Another
important study on nanostructured materials prior to the discovery of the Fe-Si-
B-Nb-Cu alloys can be found in the 1960' s. Hoffmann (1964, 1973) analyzed
the effect of grain size on soft magnetic behavior in Ni-Fe (Permalloy) films in
the nanocrystalline regime. However, the intrinsic magnetocrystalline
anisotropy in his Ni-Fe films was small enough to realize good magnetic
softness. On the other hand, recent studies on nanocrystalline soft magnetic
alloys have involved attempting to reduce the effect of large
magnetocrystalline anisotropy in Fe-rich alloys by means of grain refinement.
Consequently, the rationale behind recent work on nanocrystalline soft
magnetic alloys appears to be distinct from that of the early studies.

9. 2 Origin of Magnetic Softness-Random


Magnetocrystalline Anisotropy

9.2.1
9. 2. 1 Magnetic Anisotropies and Magnetic Softness

Good soft magnetic characteristics are obtained when both the domain wall
motion and domain rotation occur with a small change in the energy of the
system (the system here means the walls and/or the domains that directly
contribute to the technical magnetization process). One factor that may
strongly affect the energy of the system is magnetic anisotropy. Important
magnetic anisotropies in practical magnetic materials include:
342 Kiyonori Suzuki

(1) Magnetocrystalline
Magnetocrystall ine anisotropy,
(2) Shape anisotropy,
(3) Anisotropy induced by: Magnetic annealing and stress.
Among these anisotropies, the magnetocrystall ine anisotropy is the only
intrinsic anisotropy which reflects the symmetry of the crystal structure. In
addition, the stress-induced anisotropy originates from magnetostriction,
which itself is one of the intrinsic magnetic characteristics of materials.
The simplest mechanism for the technical magnetization process in
polycrystalline materials may be found in an assembly of non-interacting single
domain particles reversing their magnetization by rotation. This hypothetical
system is known as the Stoner-Wohlfarth particles (Stoner and Wohlfarth,
1948), and the total energy of the system is approximated simply by the sum
of the magnetic anisotropy energy and the potential energy due to the
magnetostatic effect. The coercivity (H c) of such an assembly of particles is
given by

Hc = P cM
Kc (9.1)
s

where K c is the magnetocrystall


magnetocrystalline ine anisotropy constant. The pre-factor Pc
depends on the anisotropy symmetry of K c and is o. O. 96 and O. 64 for uniaxial
and cubic crystals, respectively (Bozorth, 1951). Although explicit Pc values
are available under these restricted conditions (i. e., the Stoner-Wohlfarth
model), this simple linear relationship between the coercivity and the
magnetocrystalline anisotropy constant is also applicable to the technical
magnetization process by the domain wall motion (Chikazumi, 1964). The
coercivity due to hindrances to the wall motion can be related to the spatial
fluctuation amplitude of the anisotropy energy K) and the wavelength of the
anisotropy fluctuation (i\ a)'
a), i. e. ,

(9.2)

where 8 6 is the domain wall width. The origins of <K) in practical materials
may include the stress-induced anisotropy as well as the intrinsic
magnetocrystall
magnetocrystallineine anisotropy. Hence, the residual stress (a)
(0") and the
magnetostriction are also significant in this discussion. Provided that the
magnetostriction is isotropic and 0"a is uniaxial, the magnetoelastic energy
CE
(Eel)
el ) is given by

22~'1
E elel --- 3/lsO"SIn
22
3/lsasln a ex (9.3)

where i\ s is the saturation magnetostriction, and aex denotes the angle between
a. On the other hand, the magnetocrystalline anisotropy energy (E K )
M s and 0".
reflects the angle between M s and the easy axis (8), namely,
(9.4)
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 343

where K u is the induced uniaxial anisotropy. Therefore, the interference of


these two energy terms depends on the angle between u and the easy
direction (e - a). Since the magnetoelastic effect in actual materials is
attributed to the residual internal stress which varies its direction, sign and
magnitude in the sample, the EelE., in practical materials fluctuates randomly in
the sample. Consequently, E E., are usually incoherent and, hence, the
E K and Eel
amplitude of the total anisotropy energy for practical materials should be

(9.5)

Therefore, it is clear that the general concept in the material design of soft
magnetic materials is to minimize both the magnetocrystalline anisotropy
constant and magnetostriction. In fact, excellent soft magnetic materials, such
as Ni-Fe (Permalloy) and Fe-AI-Si Fe-Al-Si (Sendust) alloys, satisfy K 11 =0,
= 0, A 100 =
=00
and AliIAll1 =0 (Al00
(A 100 and AliI
All1 are the magnetostriction constants in the [100J and
[111 J directions
[lllJ directions,, respectively).
Table 9. 1 summarizes various approaches for reducing the
magnetocrystalline anisotropy energy (E K ) and magnetoelastic energy (Eel) (E., )
in bulk soft magnetic materials. In general, a high saturation induction (B, (B s =
o M s ) is expected for alloys having high Fe contents, and, hence, a small
PoM,)
J.1
amount of alloying elements is preferable for magnetic core materials.
However, in conventional crystall ine alloys, where both K K c and A A,s are
controlled by alloy composition, an addition of a relatively large amount of
alloying elements is required to obtain small E E., in Fe-based systems.
E KK or Eel
This usually leads to a low saturation induction below about 1 T in the
conventional crystalline
crystall ine soft magnetic alloys with sufficient magnetic softness.
On the other hand, small E KK in amorphous or nanocrystalline
nanocrystall ine materials is due
to their exceptionally short structural correlation lengths, and, hence, the
amount of nonmagnetic alloying elements in these materials can be smaller
than that of the conventional crystalline soft magnetic alloys.

Table 9. 1 Various approaches to the reductions of magnetocrystalline anisotropy


K ) and magnetoelastic energy (E. , ) in soft magnetic materials.
energy (E K)
Materials Preparation method EK-O
EK--O E.,-O
E.,--O

Conventional Conventional melting Composition control Composition control


crystalline alloys and casting
(Ni-Fe, Fe-AI-Si.
(Ni-Fe. Fe-AI-Si, etc.)
Amorphous alloys Melt spinning Amorphization Composition control
(Fe-Si-B.
(Fe-Si-B, Co-Fe-Si-B) ((amorphization)
amorphization)
Nanocrystalline alloys Melt spinning Nanocrystallization Nanocrystall ization
Nanocrystallization
(Fe-Si-B-Nb-Cu,
(Fe-Si-B-Nb-Cu. (amorphization) and and composition
Fe-Zr-B, etc. ) nanocrystallization control
344 Kiyonori Suzuki

Although, amorphous alloys are virtually free from magnetocrystalline


anisotropy, their magnetostriction can only be controlled by alloy composition.
In fact, amorphous magnetic alloys with As=O As = 0 are Iimited
limited to Co-rich
compositions where the saturation induction is below 1 T. Furthermore, it is
known that As in Fe-based amorphous alloys varies as the square of the
spontaneous magnetization (0' Handley, 1977; Ito et ai., al., 1980) (Fig. 9.2)
and hence, the combination of small AsA s and high M ss has not been achieved in
amorphous materials. On the contrary, for nanocrystalline structures the local
magnetostriction is averaged out over many grains within the exchange-
coupled volume and as a result, the net value of saturation magnetostriction in
Fe-based nanocrystalline alloys becomes
2 3
<A s > == "5
<As> 5A100
A 100 + "5
5A111
A 111 (9.6)

where A 100 and A 111 are the magnetostriction constants in the [100] and [111 ]
directions, respectively. Although the magnitude of A100 A 100 or AliI
A 111 in Fe-rich
alloys is as high as 10- 5 , the signs of these two constants are opposite (A (A 100
100
= + 20.7 X 10- 6 and AliI 20. 2 X 10- 6 for ex-Fe) (Lee, 1955), and,
Alii = - 20.2

hence, small As is realizable in Fe-rich compositions. In nanocrystalline


systems, zero-magnetostrictive alloys have been obtained with high saturation
magnetization values of 1.2
1. 2 to 1.7 T. It is therefore reasonable to suggest that
the approach for reducing E K and Eel E 01 in Fe-based nanocrystall ine alloys is
currently the most advanced concept of material design in soft magnetic
materials.
40
Fe-based am,orPh~
am?orPh~
s
(;('=M
(A"ocMs)
-) " " ..:... .
" ".
.."

Saturation magnetization, M s (T)

Figure 9.2 Relationship between saturation magnetostriction (As)


CAs) and saturation
magnetization (MC M,)
s ) for various amorphous (Ito
CIto et ai.,
aI., 1980), nanocrystalline
(Yoshizawa et ai.,
Fe-Si-B-Nb-Cu CYoshizawa aI., 1988, Yoshizawa and Yamauchi, 1989) and
nanocrystalline Fe(Co)-Zr-B
FeCCo)-Zr-B (Suzuki et ai.,
aI., 1994a) alloys. The saturation magnetization
of the Fe-based amorphous alloys in Ito et al. (1980) was given in emul g and the plots
4rrM s values with an assumed density 7.2 g/cm 3 .
here are the 41TM,
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 345

9.2.2 Random Magnetocrystalline Anisotropy


9.2.2.1
9.2.2. 1 Random Anisotropy Model (RAM) in Single-phase Systems
As we discussed in the previous section, good magnetic softness is brought
about by the small amp Iitude of the magnetic anisotropy energy
amplitude K .
( K) ). In this
section we discuss the effect of the grain size (0)
(D) on the amplitude of the
magnetocrystalline anisotropy energy (E K ) . The key to understanding the
reduction of the amplitude of E K in nanocrystalline materials is to realize the
relationship between the magnetic and structural correlation lengths. An
important parameter, known as the natural exchange correlation length (L ~x) ,
is determined by the competition between the anisotropy and exchange energy
terms and is defined as

(9.7)

where A is the exchange stiffness constant. The parameter cp reflects both the
symmetry of K cc and the spin rotation angle over L~xL ~x and is typically of the
order of unity. This natural exchange correlation length L ~x corresponds to the
critical grain size below which the amplitude of E K starts to decrease with
grain refinement
refinement. Figure 9.3 describes the reduction process of the amplitude
of E KK in nanocrystalline materials. A uniaxial system (K ce = K u)
u ) with a unit
volume was chosen here for simplicity of the argument
argument. The angle dependence

~ Kr
OJ

"
OJ

it
e

Figure 9.3 Change in the magnetocrystalline anisotropy energy (E K) K ) for conventional


(D > L ~x) and nanocrystall
(0 nanocrystalline (D L ~x) materials (for uniaxial anisotropy symmetry) ,
ine (0
E K ( K c
showing that the amplitude of EK(K c ) is suppressed to smaller than K c 0 L~x.
c for D

The total anisotropy energy per unit volume is independent of the grain size (i. e.,e , Kc
remains constant) and the symmetry of K c is preserved in (K c) .
346 Kiyonori Suzuki

of E KK for a large grain (0


CD>> L ~x) is described by Eq. 9. 4 in the previous
K is equal to K c'
section and the amplitude of E K c . On the other hand, E K
K for 0

L ~x is given by the sum of the contributions from a number of grains C (E~),


E~ ) ,
i.e. ,

(9.8)
C9.8)

where N is the number of the magnetically correlated grains. An important


physical assumption in Eq. (9.
C9. 8) is that the anisotropy energy of each grain
inside the exchange correlation length is mutually transmitted to each other via
the exchange interaction. Here, the magnetocrystall ine anisotropy energy of
CED can be described by taking into account the angle between its
each grain (ED
=
easy axis and the reference axis at 8 = 0 (Y
Cy;)
i) :

. K 2
E K = NCsin
E~= NCsinC8-Y;).
(8 - Yi)' (9.9)
C9.9)

Provided that the grains are oriented at random, the mean square amplitude of
the anisotropy energy is composed of the quadratic contributions of the random
local anisotropies from different grains. Hence, the ampIitude of the
magnetocrystalline anisotropy energy in Eq.C9.8),
Eq. (9.8), i.e.,
i. e., the random
magnetocrystalline anisotropy C ,
(KK cc), is obtained by the following random-
walk consideration:

<K =
(K c >) = =
JNCED 2 =
IN(ED !;N
:Jir (9.10)
C9.10)

where EK ~ is the ampIitude


amplitude of the local anisotropy energy. Since this random
magnetocrystalline anisotropy corresponds to the spatial fluctuation of
anisotropy energy in the sample, which dominates the domain wall motion
(Eq.
CEq. (9. 2) ), a small H
C9. 2)), Hec is obtained even for materials with large intrinsic K c .
It is worth remarking here that the intrinsic K K c in each grain remains the bulk
value and only the amplitude of the magnetocrystalline anisotropy energy is
reduced. This is readily understood from the fact (Fig.CFig. 9.3)
9. 3) that the integrated
area of the E K-8K -8 curve is independent of the grain size. In addition, the
anisotropy symmetry of K c is preserved in the fluctuating part of the E K-8
curve for 0 L L ~x .
A significant aspect in the random anisotropy model (RAM) CRAM) is that the
anisotropy energy in Eq. (9.7)
C9. 7) is also suppressed to (K c > in the range 0
< K c)
L ~x and consequently, the exchange correlation length itself starts to expand.
Naturally, this expansion results in a further reduction of (K < K c >,
), and the

averaging process of E KK is accelerated. The analytical solution for (K <K c)


c > under

the framework of the random anisotropy model for the condition 0 L L~x
~x was
C1989, 1990a, 1995), and we follow his argument below.
given by Herzer (1989,
The number of grains in a magnetically coupled volume which, in bulk-
form systems, is
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 347

N (cixff
= (L~x (9. 11)

where the effective exchange correlation length L ex


ex in the nanocrystalline
nanocrystall ine
regime is given by

(9. 12)
(9.12)

where <(KK > is the amplitude


ampIitude of the effective anisotropy energy. Provided that
the random magnetocrystalline anisotropy is the dominant anisotropy in the
system and the effective anisotropy constant is approximated by (K ec > (i. e. ,
(K>~(Kc'
< K >~ <K e, the length of L
Lex
ex can then be determined self-consistently using
Eqs. (9. 10) to (9.12), yielding

(9. 13)
(9.13)

which leads to the well-known relation:

(9. 14)

In this argument, Herzer assumed <p = 1 rad (cp


cp = (<p == ./473
A73 rad for L exex in the
model of Alben et a!. (1978)). The origin of these <p
al. (1978. cp values was not discussed
in detail in their models, and hence the absolute value of (K <K > is not entirely
free from uncertainty. However, the significance of Herzer's approach lies in
the derived power laws: 0 6 , K~ and A -3. It is important to restate that the
above seal ing-I ike argument is only val
scaling-like valid
id under the condition (K >~ ( K c >.
(K>~<Ke>'

9.2.2.2 RAM Under the Influences of Induced Anisotropies


The effective anisotropy in the nanocrystalline
nanocrystall ine materials may have
contributions from induced anisotropies other than the random
magnetocrystalline anisotropy and hence the effective anisotropy constant in
actual systems is more correctly as.

(9. 15)

This term comprises at least the contribution from the magnetoelastic energy
as in Eq. (9. 3). Consequently, if the K K u term in Eq. (9. 15) is significant
relative to (K ec > (e. g. samples with large magnetostriction and small grain
<K > value is governed by K u' In such a case, one may no longer
sizes), the (K
see the effect of grain refinement on the coercivity variation. More
importantly, the critical condition <K > ~ (K
(K <K ec > obviously does not hold and
<K > alters the re-normalization
(hence) the contribution of K u to (K re-normal ization process of
Lex and both Eqs. (9. 13) and (9. 14) change. Therefore, an exact application
(9.14)
of the 0 6 , K~ and A -3 scaling-like rules should be strictly limited to those
samples whose magnetization process is governed by the random
348 Kiyonori Suzuki

magnetocrystall ine anisotropy c >..


an isotropy < K c)
Herzer (199Gb) and Suzuki et al. (1998a) re-analyzed the random
magnetocrystall ine anisotropy by taking into account the effect of coherent
magnetocrystalline
uniaxial anisotropies on the effective exchange length. In the original random
anisotropy model (Eq. (9. 14, the 0 6 dependence of H c is derived from the
following relation:

(9 16)

On the other hand, if there is an additional uniaxial anisotropy contribution


determining Lex' then the effective anisotropy constant in such a sample
becomes

(9.17>
(9.17)

Lex """ VA / <K)


where the relationship Lex::::::::: K > must be maintained (q> ::::::::: 1 is assumed
(cp """
here for simplicity). Although this equation cannot be solved analytically for
<K),
<K >, a limiting condition K uu <K <K c
c>
) enables us to arrive at

(9.18)

In the case of K uu c >, Eq. (9. 17) simply reduces to the result of the
<K c),
original RAM (Eq. (9.14.
(9. 14.
As we discussed in the previous section the coercivity due to hindrances
to the wall motion (Eq. (9.2 is a function of the wavelength of the anisotropy
CA a) as well as the ampl
fluctuation (i\ itude of the anisotropy energy K .
amplitude >). If
the wavelength of the uniaxial anisotropy component in Eq. (9. 18) is
considerably larger than the domain wall width (i. e. , i\A a a Lex),
Lex)' the second
Eq. (9. 18), which corresponds to <K
part of Eq.(9.18), < K cc >,
)' ultimately determines the
coercivity. Thus, the grain size dependence of H c changes from the well-
known 0 6 law to a 0 3 variation when the coherent uniaxial anisotropies
dominate over the random magnetocrystalline anisotropy (i. e. , when both the
a L exex and K u
conditions i\Aa u < K cc)> are satisfied). It is worth noting that
similar analysis, which is essentially based upon scaling arguments and
statistical considerations, can already be found in the classical paper of Alben
et al. (1978) who explicitly considered the situation of domain wall
displacements when a coherent anisotropy adds to local random anisotropies.

9.2.2.3 RAM in Two-Phase Systems


Another possible source of discrepancy between the single-phase model and
experiments is the fact that the models discussed above assume single-phase
materials while the nanocrystalline soft magnetic materials prepared by
primary crystallization of amorphous precursors contain a residual amorphous
phase in the intergranular region. The extension of the random anisotropy
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 349

model to two-phase systems was carried out by many authors (Herzer, 1995;
Hernando et aI., 1995, 1998; Suzuk
Suzukii and Cadogan, 1998b; Loffler et al. ,
1999). Two extended models to be discussed in this section are (1) RAM in
multi-phase systems with single exchange stiffness constant and (2) RAM in
amorphous/
amorphousl crystalline two-phase systems with two distinct exchange stiffness
constants.
An important aspect of multi-phase systems is the presence of more than
two energy densities for the local K c' Herzer (1995) analyzed the random-
walk process of such a system by taking into account quadratic contributions of
<K > in this multi-phase
the local magnetocrystalline anisotropy constants and (K)
system is given by

(9.19)

where 0;0 i and K;K i are the mean grain size and the local magnetocrystalline
anisotropy constant of ith-phase, respectively. The number of grains in the
coupled volume for the ith-phase can be given by the relation N N;i = V; exl/
Vi (L ex

o i) 3, where V;
0;)3, < K > is
V i is the volume fraction of the ith-phase. Hence, (K)
determined self-consistently using Eqs. (9. 12) and (9. 19). This leads to the
< K > for multi-phase nanocrystall ine systems:
following (K)

<K >
(K) ~ (~V~3~~7f
(~ V ~3~~ 7 ) 2 (9.20)
I

Provided that the effective local anisotropy constant of the residual amorphous
bccl amorphous two-phase nanocrystalline materials is negligibly
phase in the bee/amorphous
small relative to K 1 of the bee-Fe nanocrystall ites, Eq. (9. 20) is
approximated by

<K>~(1-Vam )2
6
Ki0
A3 (9.21)

where V am is the volume fraction of the residual amorphous phase. This result
physically means that the magnetocrystalline anisotropy energy is diluted by
the volume of the amorphous component. Obviously, this dilution effect only
1>' and the total anisotropy energy may increase with V am if the
< K 1)'
applies to (K
induced anisotropies in the amorphous phase are considerable. It should be
noted that the original 0 6 , Ki
Kj and A -3 laws are maintained in this two-phase
model at a constant Vam value.
In Herzer's two-phase random anisotropy model above, the angle
between the nearest spin-spin pair within the coupled volume is assumed to be
constant for both the nanocrystalline and intergranular regions (typically
J A 1<
d /I V K > rad, where d is the spin-spin distance). However, if the
/ (K)
exchange stiffness in the intergranular amorphous phase is lower than that of
the crystalline phase, the spin rotation should become more rapid within the
amorphous part compared with that in the crystalline part with the higher
350 Kiyonori Suzuki

exchange stiffness. Hence, two distinct angles for the spin-spin pairs in the
crystalline and amorphous regions were considered in the two-phase RAM
model of Suzuki and Cadogan (1998b), in order to include the effect of the
exchange stiffness in the intergranular phase on the exchange correlation
length.
In a two-phase system where the exchange stiffness constants of the
crystalline and amorphous phases are A cr and A am , respectively, and the
thickness of the intergranular amorphous phase is 1\ (F ig. 9. 4), the spin
rotation over each crystallite-amorphous coupling pair with the length 0 + 1\
would be

(9.22)

where 1\ is given by assuming a simple cubic geometry for grains, i. e. ,

(9.23)
Since the total spin rotation over the exchange correlation length is defined by
Ip, the number of grains within the entire coupling chain over L ex is
<p,

nn =
= .
~ == - - - - - - ----LIp
' < p - - - - _- (9.24)
<pO
mO 0 1\
't' ~::::;==~ +
+-----;::;:.:::::::;::~
-----;:;:.:::::::;::~
JAcr/<K> JAam/<K>
Crystalline

,---I'_D_--'''' .. n _ _ --'liD
HIll kdII!H,/~/////
Figure 9. 4 Schematic illustration describing the grain size (D) and the thickness of the
(/I) in the two-phase RAM (Eq. (9. 29) .
intergranular residual amorphous phase (1\)

Furthermore, the magnetically coupled volume in this system is equal to


L 33 =
L
ex
= 1-V
(nO )3
(nD)3
l-V
(9.25)
(9.25)
am

and hence L ex in the system is given by


Lex
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 351

=
Lex = __ 1
1'----------,--, _-----,=---_ _tpO
cp<..::O"-----_---:/\_ _
--'----------,--1 -------:=-----'-='-------,------ (9.26)
(l-
(1 -V
V am ))"3
"3 0
0 /\
+ -----;==;==:=
VAcr/<K)
/Acr/<K) VAam/<K)
/Aam/<K)
where <K) is given by

<K)~<Kl)=(l-V
<K)"",,<K 1 )=(1-V am )~
IN
Jf\i (9.27)

and the total number of grains in the coupled volume N is

N = (1 - Vam ) ~~;. (9.28)

The exchange correlation length can be self-consistently determined using


Eqs. (9.26) to (9.28) and one arrives at
Eqs.(9.26)

1 cp 4 0
tp
Lex~
Lex ~ 7 (9.29)
)"3 K~(OI
(1 - Vam )3 K~(OI JA: + /\1
J""A: ~)4

and hence <K) in this two-phase model is given by

<K) "'-' 1
""" 6(1 - Vam ) K 4(~+~)6
1 /1\ ~
4 (9.30a)
cp VAcr vAam
or, equivalently,

It is worth noting that this final result of the extended two-phase model
reproduces the same result of Herzer's two-phase model (Eq. (9. 21 under
= Aam
A cr =
the condition Acr A am and for V am = 0 it reduces to the original single-phase
am =
nanocrystall ine random anisotropy model (Eq. (9. 14. Important features in
the result of the above analysis are CD the well-known 0 6 dependence is
preserved and CV A ;m3 under the condition Aam
~ <K) varies as A;;-m A cr' physically
A am Acr'
because the spin rotation within the exchange correlation length occurs mostly
in the amorphous region.

9.3
9. 3 Nanostructure-Magnetic Properties Relationships

9.3. 1
9.3.1 Grain Size and Magnetic Softness
The intrinsic material parameters that dominate the amplitude of the
magnetocrystalline anisotropy energy K in nanostructured magnetic
352 Kiyonori Suzuki

materials include the grain size (0), the magnetocrystalline anisotropy


constant (K e) and the exchange stiffness constant (A). Among the effects of
these dominant parameters, the relationship between the grain size and the
coercivity has been studied extensively in many alloy systems. Figures 9. 5a, b
show the relationship between the mean grain size and coercivity in various
Fe-based nanocrystalline soft magnetic alloys (Fujii (Fuj ii et aI., 1991; Suzuki et
aI., 1991 a, 1996; Yoshizawa, 1993; Muller and Mattern, 1994; Herzer,
1995). As shown in Fig.9.5a,
Fig. 9. 5a, He of the nanocrystalline Fe-Si-B-M-Cu (M=(M =
IVa to Via metal) alloys follows the predicted 0 6 dependence in a 0 range
below LL~x
~x (
(--
' " 30 to 40 nm for this alloy system) although the plots deviate from

the predicted 0 6 law in the range below He -- 1 Aim where the effects of
He'"
induced anisotropies (e. g. , magneto-elastic and field induced anisotropies) to
He seems to be significant relative to those of the random magnetocrystalline
anisotropy. Consequently, the fluctuation amplitude of the total anisotropy

102
He =D6 0
0

0
0
10 1 =Ja2+(bD6 )2
H e=)a
~
E
3-
~
0::0
:I:;0
0
Fe-Si-B-M-Cu
0 (M=IVa to Via metal)
100
10 0 oo I
<eo
~D (IJ
00 0/
0/ 00 o (Yoshizawa)
/
I/ (Herzer)
I/ 0
o (MUller
(MUlier and Mattern)
H e =D6I/ /
10-1
10 1 10 2
2

Mean grain size, D


D(nm)
(nm)
(a)
102
Hec =D3

10 1
10'

I3-:t:"E
::r::"
100
10 Fe78PxCI 8-xGe3Si OsCuo Ss
o Fe78PxC'8-xGe3Sio
(x=9 to 16)(Fujii et al.)
aL)
Fe9lZr7B2(Suzuki
Fe9,Zr7B2(Suzuki et al.)
aL)
Fe90Zr7B2CUl(SUzuki et al.)
oo Fe90Zr7B2Cu,(Suzuki aL)
10-'
10- 1
10 1 102
Mean grain size, D(nm)
(b)

Figure 9. 5 Coercivity (He) versus grain size (D) for (a) Fe-Si-B-M-Cu (M = IVa to Via
metal) (Yoshizawa, 1993; Muller and Mattern, 1994; Herzer, 1995) and (b) Fe-Zr-
B(-Cu) (Suzuki et ai.,
al. , 1991a, 1996) and Fe-P-C-Ge-Si-Cu (Fujii et al. ,1991)
,1991> alloys.
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 353

energy should be better described by Eq. (9. 15). The He variation in


Fig. 9. 5a is, therefore, fitted by assuming contributions from both CD the
spatial fluctuations of induced anisotropies and CV < K I1 > to He ( i. e., He =
<Z) (K
2
va + (b0 )2, where a ~
6
""'" 1 Aim is the contribution originating from the
induced anisotropies). This clearly indicates that < < K 1 > in the nanocrystalline
Fe-Si-B-M-Cu alloys can be described by the random anisotropy model.
Unlike the Fe-Si-B-M-Cu alloys, He of the nanocrystalline Fe-Zr-B (-Cu)
and Fe-P-C-Ge-Si-Cu alloys shown in Fig. 9. 5b appears to follow a simpler 0-
power law with an exponent of 3 and no clear deviation of the plots from the
power law is seen. This indicates that He in these alloys is chiefly governed by
<K 1> which varies as 0 3 , suggesting the exchange correlation length under
the influence of coherent uniaxial anisotropies (i. e. , LLex ex under the conditions

K u <K e >) .
i\ a L ex and K
A .
The presence of uniaxial anisotropies in Fe-based magnetic materials can
be reflected in their hysteresis loops (Tonge and Wohlfarth, 1958). Figure 9.6 9. 6
shows typical hysteresis loops of nanocrystalline samples exhibiting the 0 3 and
0 6 behaviors. The remanence to saturation ratio (Brl (BrIB B ss )) of the
FenSiluBsNb15Cul
FenSil25BsNb15CuI sample, whose He and 0 values are in the 0 6
dependence regime in Fig. 9. 58,5a, is about O. 95, clearly enhanced above the
theoretical value for cubic Stoner-Wohlfarth particles (0.83). This is a
characteristic feature of the interplay of the exchange interaction and random
anisotropies in the small grain size regime. Yet at smallest grain sizes, when
the random anisotropies are sufficiently averaged out, more long-range
uniaxial anisotropies take over the control and the remanence ratio decreases
to values around 0.5. Thus, the Fe9,Zr7B2
Fe91 Zr7 B2 sample shown in Fig.9.6,
Fig. 9.6, with its
reduced Brl B s of only about 0.45, is a typical example for the presence of
more long-range uniaxial anisotropies larger than the averaged magneto-
crystalline anisotropy <K 1>'
I>' The fact that He in these Fe-Zr-B( -Cu) alloys still
1.0
FenSi 12SBgNb15CUI
FenSil2sBsNbuCUI
BrlEs"'" 0.95
BrlBs"" I / / /---
/--
He
H c :21 Aim I / I/
0.5 D:25
D: 25 nm {/ I
IIII I/
g I I I/
~ 0o Fe91Zr7Bg --:1-1---
--:1--1---
Fe91Zr7Bs I I/
BrlBs ""
"'" 0.45 I 1/
-0.5
--0.5
Hec :4.2A/m
H :4.2 Aim
D:12nm
/ )
__ / / / / II
I
-1.0b:;;!~~-~'=~-==-=-;::::==~==:::::::'_--=,::-
-1.0 -:,:::-_
b:;;;;!;~~===~c::::::::==::1::=:::::=----~-------:';::---
-40 -20 0 20 40
H(Alm)
H(A/m)

Figure 9.6
9. 6 DC-hysteresis loops of nanocrystalline Fen Si"o
Si ,z. s5 B
8 8 Nb 15 Cu,
CUI and Fe91 Zr782
Zr7 Bz
alloys annealed at 813 K for 3.
3.66 ks and 923 K for 60 s, respectively (Herzer, 1996;
Suzuki et ai.,
aI., 1998a).
354 Kiyonori Suzuki

sensitively depends on the grain size indicates that He is still governed by


<K 1 >,
>, i. e. the spatial fluctuations of the induced anisotropies are smaller than
<K 1>
I> and do not dominate He like in the case of the Fe-Si-B-M-Cu alloys at
small grain sizes (20 to 30 nm). However, the presence of these quash-
coherent induced anisotropies obviously changes the grain size dependence
significantly to a 0 33
law. Such anisotropies may be induced by the domain
structure during annealing or magneto-elastic interactions.
The reduction of < K e > in nanostructured systems is due to the expansion
<K
of Lex. Naturally, the Lex cannot exceed the actual sample size. size, and, hence,
the sample dimension could be another significant parameter in determining the
coercivity of nanostructured materials. By replacing the actual material
nanocrystall ine soft magnetic alloys, 0 = 10 nm,
parameters of Fe-based nanocrystalline
= 10 kJ/m 3 and A =
K 11= = 10- 11 J/m 2 , into Eq. (9.13),
(9. 13), the Lex in a typical
nanocrystalline sample is estimated to be '" -- 1 ~m, this estimation is verified by
the domain observation of nanocrystalline Fe735Si135BgNb3Cul
Fe735Sil35BgNb3Cul (Schafer et al. ,
1991). Since the typical thickness of melt-spun ribbons is about 20 - 50 ~m,
the L Lex
ex in melt-spun nanocrystalline samples is virtually unlimited in any
directions. On the contrary, the L ex in thin film-form samples may be limited in
Lex
the direction of the film thickness as the thickness of sputtered films is usually
in the order of ~m. In such a case, the L ex can only be expanded in
2-dimension. The random magnetocrystalline anisotropy in am-dimensional
system is obtained by simply substituting Eq. (9.11) (9. 11) with

N = (L~xf
(L~x( (9.31)

This substitution leads to the following solution for <


< K >:
2m

<K> ~ Ke[~J-m.
Ke[~J4-m. (9.32)

This general solution for < K > predicts that the coercivity in thin-film
nanocrystalline materials (m = = 2) varies as the square of the grain size. In
Fig. 9. 7 we show the grain size dependence of nanocrystalline FeBI 4 TaUClO.3
TaUC103
films prepared by sputtering (Hasegawa, 1993b). A D-power O-power dependence
with an exponent of 2.5 is derived from least-squares fitting of the data points.
This exponent is substantially lower than the exponent for 3-dimensional
systems (6) and closer to the theoretical exponent for 2-dimensional systems
(2), confirming the significant effect of the sample dimension on the magnetic
nanocrystalline
softness of nanocrystall ine systems.
The relationship between the grain size and the coercivity in thin-film
nanocrystall ine materials was analyzed (Hoffmann, 1973) before the
development of the random anisotropy model. The spatial fluctuation of the
anisotropy energy in this model is also derived from the assumption that the
contribution of the magnetocrystalline anisotropy follows the random-walk
(Eq. (9. 10)), although the shape of the magnetically coupled region in
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 355

1000

10

10 100
Mean grain size,D(nm)
size,D (nm)

Figure 9.7 Coercivity (He) versus mean grain size (D) for nanocrystalline
FeS14 TaUC103 films (Hasegawa, 1993b).
FeS1.4 Ta8.3C103

Hoffmann's theory is ellipsoidal due to the consideration of the dipole-dipole


interaction. The coercivity under the framework of this so-called ripple theory
(Hoffmann and Fujii, 1993) scales as

(9.33)

where K u is the induced anisotropy, and t denotes the film thickness. Hence,
it appears that both the random anisotropy model and the ripple theory predict
the 0 2 dependence for the coercivity in 2-dimensional systems. The ripple
theory has only been developed for 1.5- or 2-dimensional systems to date and
(9. 33) is strictly limited to the
it should be noted that the application of Eq. (9.33)
systems where the magnetization is fully aligned in plane and the induced
anisotropy is relevant to the domain wall width (i. e. , the systems where K u
<KK c)
< c > is valid).
valid) .
We have so far tested the applicability of the random anisotropy model to
actual nanocrystalline systems based on the scaling property of the coercivity
in terms of O. However, the validity of the absolute < K K>> value in RAM is
still open to question. We may argue this crucial point by analyzing the
experimental data for the nanocrystalline Fe-Si-B-M-Cu alloys shown in
Fig. 9. 5a. As we discussed, the coercivity of nanocrystalline Fe-Si-B-M-Cu
alloys is well fitted by the following
follow ing form: H = J a0 22 + (b0
Hec = ( bO 66 ) 2. The pre-
factor b is estimated to be 5 x 10 46 A/m 7 for the nanocrystalline Fe-Si-B-M-
Cu alloys by least-squares fitting. Provided that the coercivity in the
nanocrystalline Fe-Si-B-M-Cu alloys is due to the hindrance of the wall motion
(Eq. (9.
(9 . 2)) by the random magnetocrystalline anisotropy, the pre-factor b
is calculated from
356 Kiyonori Suzuki

~ _1 CK)
b """ (K> ~ """
~ (1 - V am )2Kj
)2Ki ~ ~ """ (1 - 0.3)2(1 X 10 4 )4 X 0 5
6 Aa
0 M s i\a M 6
M S cp A 3 Aa
i\a 1.34 6 (1 X 10-11)3
10-11)3 X ..5
48
~ 1.9
""" 1. 9 X 10 (Aim?) (9.34)
cp6

where the domain wall width 0 is approximated by Lex' and the wavelength of
CK Lex ( i. e., 0
1 ) should be at least twice the length of Lex(i.e.,
( K 1> I Aa '" 0.5).
oli\a'" O. 5). On the
other hand, cp could be estimated from the critical grain size (Do), below
which the coercivity starts to follow the 0 6 dependence; Do is about 50 nm for
the nanocrystalline Fe-Si-B-M-Cu alloys (Herzer, 1997). Since this critical
grain size corresponds to the natural exchange correlation length, cp for the
nanocrystall ine Fe-Si-B-M-Cu alloys is estimated from the following relation:

= L ~x I IA7f(;
cp = IIfTK: ~ """ 0Dol IIfTK:
0 I IA7f(;

= 5 X 10-
= 8
J (1 X 10 11)
10- I .j 11) I (1 X 10 4 ) ~
""" 6.
1. 6. (9.35)
Assuming this estimation for cp in Eq. (9.34), we finally obtain 12 x 10 46 Aim?
for b, slightly larger than the experimental b value (5 x 10 46 Aim?). Given the
Aa used in the above estimation may contain a large
fact that the ratio of 0 to i\a
error, the estimation of b based on the random anisotropy model appears to
be quite realistic. It is restated here that the random anisotropy model
proposed by Herzer (cp = 1) is a semi-quantitative model where the
significance lies in the scaling-like argument. As exemplified in the above
estimation of b, the absolute value of (K 1> in RAM needs to be discussed in
CK 1)
conjunction with the analysis of the spin rotation angle cp.
Some authors estimated the macroscopic magnetic anisotropy of
Fe775Si135BgNb3CUI by analyzing the magnetization curves (Ho
nanocrystalline Fe775Si135BgNb3Cul
et al. , 1993; Varga et al. , 2000). The estimated anisotropy constants were
about two orders of magnitude larger than (K < K 1)
I > in the random anisotropy
model (typically about a few J/m 3 for Fe775Si135BgNb3CU1)
Fe775Si135BgNb3Cu,) while the
coercivity values reported by these authors were consistent with the usual
Fe775Si135BgNb3CUI alloy ('" 1 Aim). This discrepancy between
value for the Fe775Si135BgNb3Cu,
the experiments and RAM may be understood by taking into account the
influence of coherent anisotropies. As we discussed in Section 9. 2. 1, the
coercivity reflects the fluctuation amplitude
ampIitude of the anisotropy energy. Hence,
excellent magnetic softness can be obtained even with the presence of large
anisotropies as long as the anisotropies are coherent or the wavelength of the
anisotr6pies
anisotropy fluctuation is considerably longer than the exchange correlation
length. On the contrary, the coherent anisotropies are reflected in the
magnetization rotation process in high magnetic field ranges. Hence, the
anisotropy constants estimated from magnetization curves in the previous
studies could be due to the coherent anisotropies rather than owing to the
random magnetocrystalline anisotropy < 1> that is relevant to the coercivity.
K 1)
(K
Further clarification of the origin of the large macroscopic anisotropy constant
is required by looking at the symmetry of the dominant magnetic anisotropy in
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 357

the sample.

9.3.2 Intergranular Phase and Magnetic Coupling

In the temperature range below the Curie temperature of the residual


intergranular amorphous phase (n m
m
),
) , the nanocrystallites are coupled
magnetically via the exchange interaction acting over the bcc-amorphous-bcc
coupling chain. However, this coupling chain is diminished in the temperature
n
range above n m and the nanocrystalline alloys behave as an assembly of
isolated magnetic particles in which the magnetically hardest domain
configuration is expected. Consequently, we observe a significant increase in
Hec in the temperature range above ....,
H --- n m
(Herzer, 1989). This effect is a
(Herzer.
possible disadvantage of the nanocrystalline soft magnetic alloys from the
appl ication viewpoint. On the other hand, the intergranular residual amorphous
application
phase plays important roles. The presence of the residual amorphous phase is
essential to maintain the metastable thermodynamical equilibrium of the
nanostructure. In addition, the random magnetocrystalline
magnetocrystall ine anisotropy could be
di luted by the volume of the residual amorphous phase (i. e..
diluted e., the effect
(9. 21) and, hence, the presence of the residual amorphous
described in Eq. (9.21>'
phase may be favorable to magnetic softness while the magnetic decoupling in
the intergranular region is unfavorable to it. Therefore, the volume fraction and
the exchange stiffness of the residual amorphous phase are particularly
important in order to understand the magnetic softness in the two-phase
nanocrystall ine systems.
A sophisticated way of examining the relationship between the strength of
the intergranular coupling and the macroscopic magnetic softness in two-phase
fact. to look at the temperature dependence of
nanocrystall ine systems is, in fact,
n
the coercivity in the vicinity of T em . There are at least five different
parameters other than the exchange stiffness constant in the intergranular
region that may influence the effective magnetocrystall ine anisotropy (cf. the
two-phase RAM as in Eq. (9. 30b) ). However, <p
(9.30b. cp,, Vam and 0 are all
temperature independent parameters. In addition, the temperature
dependence of both K 1 and A crcr near n m
n
is negligibly small if m is well below
the Curie temperature of the crystall ine phase (Tg). Consequently, A am can
virtually be the only temperature dependent parameter in the temperature
n n
range near n m under the condition n m Tg and the coercivity should chiefly
reflect the variation of A am'
Figure 9. 8 shows the temperature dependence of the coercivity for
nanocrystalline Fe91 Zr7 8 2 annealed at 823 K for 60 s (V am
nanocrystall ine Fe91Zr7B2 O. 5 and 0 """
""" 0.5
am ""=' ""='
13 nm). The coercivity increases slowly with temperature from 5.7 AI m at
77 K to 56 Aim at room temperature and then begins to increase significantly
decoupl ing effect. This magnetic hardening effect is well
due to the magnetic decoupling
described by the extended two-phase RAM.
358 Kiyonori Suzuki

10oor----------------,1.o
1 0oor----------------,1.o

800

600 0.6
:F
:
~o . ~
':i
~
:J::"
::r:: 400
I
..
!
I
0.4

--,/
.-//
200 0.2

oOl--~I~O~O-~20~0~~3'""0:;;-0---:40:!:-;0;----5~0~
l-.-~1~0~0--~20~0~~3~0:;c0---:40:!:";0;----5;-;(0~
Temperature(K)
Temperature (K)

9. 8
Figure 9.8 Temperature dependence of the coercivity for nanocrystalline Fe91 Zr,8
Zr7822
annealed at 823 K for 60 s (Suzuki and Cadogan.
Cadogan, 1998b).

It is known that the exchange stiffness constant of materials varies as


2
Ace T Cc S
Aoc (9.36)
d
where S is the spin.
spin, and d is the lattice parameter. Hence, the exchange
stiffness should scale as the square of the spontaneous magnetization. On the
other hand, the temperature dependence of the spontaneous magnetization for
n
the residual amorphous phase (M~m) near T't{"m is expressed by

Msam ((T)
T) = M0 am
0
(1 - ~) ~
T~m
(9.37)

= O. 36 gives
where f3 is the critical exponent. It has been confirmed that f3 =
reasonable fits to thermo-magnetic curves in various nanocrystalline materials
(Herzer, 1989; Slawska-Waniewska et al., 1992). From this relation, one
can expect that A am ( T) varies near m
n
as A am (T) ce
oc m
- T)2~. Since(n
am is considerably smaller than A cr
A am cr in the temperature range near
m
and A am
am n
is virtually the only temperature dependent parameter in this temperature
range, < K ( T in the extended two-phase RAM (Eq. (9.
<K(n) 30b varies as
(9.30b
<K(T oc (nm - T)-6~.
<K(Tce(nm-n-6~. (9.38)
Hence, a critical behavior is seen in the plots of H; 1!6~
1/6~ against T (Fig. 9.9 a). The
critical temperature is 370 K. The Curie temperature of the residual amorphous
phase in this sample estimated from the temperature dependence of 57 Fe
hyperfine field (375 K) is in agreement with the critical temperature in the
H; 1!6~
1/6~ plots, indicating that the random magnetocrystalline anisotropy in the
two-phase nanocrystalline systems can be well described by Eq. (9. 30b) .
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 359

0.5 1.0

,"
x
z/x""xx pxxxx
x""x.x~vXJtxx ~x
' XX ~
~'" ~~ -< 'x",,~xx'i<\"~
x~~:YA ~~
"'-
"'"
'>e
f>
-~
~
0.4 ?""'" '"
?"",x,\
",xx,,
$i,Jf
.....,.,. ' -
x
.~xx,
"
xx X;i x xiX<
x x ~
NIx""
x x 1!< 0.8

E 0.3 0.6
::;:
~ ;f
~.
~
'>e
f>
0
0.2 0.4
0.4 ::f
::i
'
::t;
:t:
0.1 0.2

0 0
Temperature (K)
(a)

- - - - 1 - _ ...... _
20 ~'I-,
't-~
~
"i--..
'i,
~, am
am
Tec (362 K, from
15
E 'i
'i,>.-
'........
... thermal scan)
oJ~ 10
"-
""-\
\ ~~
\
(Bhf(0=20.8 T
(Bht1.0=20.8 \
5 \
TeamTI =375 K
Tt \
\
0 100 200 300 400 500
Temperature (K)
(b)

Figure 9. 9 Plots of (a) the coercivity to the power - 1/6/3 (/3 = 0.36) and (b) the mean
1/6,8 (,8
hyperfine-field versus temperature for nanocrystalline Fe91 Zr, 6 2 annealed at 823 K for 60 s
Zr7 8
(Suzuki and Cadogan, 1998b).

Figure 9. 10 shows some simulations of < < K 1 >, in accordance with


Eq. (9. 30b),
30b) , for a typical two-phase nanocrystalline system (0 = = 10 nm,
K 11 = 10- 1111 J/m2 and cp=
= 10 4 J/m3 , A cr == 1O- q> = 1) for various values of A am and VV am .
These simulation results indicate that <K 1 > decreases with Vam V am when A am ~ """
A cr whereas it increases with Vam Vam when AamA cr . Hence, the optimum volume
fraction of the residual amorphous phase to realize good magnetic softness
should vary depending on alloy systems. Naturally, these simulations are only
for the random magnetocrystalline anisotropy K 1 while the magnetic
softness in practical materials may also be governed by induced anisotropies.
Hence, attention should be paid to the anisotropy symmetry. As mentioned in
Section 9. 3. 1, the dominant anisotropy in an Fe-based system can be
determined by looking at the remanence to saturation ratio and one may
confirm the significance of induced anisotropies.
Another interesting aspect of these simulations is that the <K 1 > shows a
peak when A am is considerably smaller than A cr' Hence, H c of nanocrystall ine
alloys with small A am may increase at the initial stage of nanocrystallization.
Such a magnetic hardening effect has been confirmed for nanocrystalline nanocrystall ine
360 Kiyonori Suzuki

100
D=10nm
D=IOnm
SO K 1I =104 J/m 3
I 0- 11 Jlin
A bcc = 10- J/m 2

!
~

60

o 0.2 0.4

Figure 9. 10 Change in calculated (K <K 1>


I> at various A am as a function of volume fraction of
the residual amorphous phase ((V
Vam )) ..

Fes6Zr7B6Cu, (Suzuki et al. , 1991c).

9.3.3 Application-Oriented Magnetic Properties


As we discussed in Section 9. 2. 1, exceptionally small values of <i\ < As>
s > have

given the nanocrystalline soft magnetic materials the most advanced family of
magnetic core materials. The small values of <As> in Fe-based nanocrystall ine
soft magnetic materials is, in fact, the result of the balance between the
m
positive magnetostriction in the residual amorphous phase (A: (A~m)) and the
negative magnetostriction in the bee-Fe phase (A ~cC).
~CC). In Fig. 9. 11 we show
the relationship between V am and As for nanocrystalline Fe91Zr7B2 as an
example (Suzuki et al., 1996). Since the volume fraction of the residual
amorphous phase decreases with increasing annealing
anneal ing temperature, the net
magnetostriction in two-phase nanocrystalline materials can be adjusted by
changing annealing conditions. This is one of the great advantages of the two-
phase nanocrystall ine materials from the viewpoint of alloy development.
nanocrystalline
2
m
A-.= V.m
As= ~m +(1-
v"mx: Vam)X;'
+( I-Vam)x;r
X;r=(-7.60.S)X 10-6
X;'=(-7.60.S)x
~ x:m=(S.S 1.4)X 10-6
t 0
x
::r -]
-I

-2

-3'-:----::-':-,,---------,.~--~-:-------,._L.,._--~.
-3L..,----::-':-:,---~-:-::----=-~--~c:__-~.
0.30 0.35 0.40 0.45 0.50 0.55
Vam
Figure 9. 11 Relation between the volume fraction of the residual amorphous phase ( Vam )
and saturation magnetostriction Os)
Us) for nanocrystalline Fe91 Zr7 8 2 (Suzuki et ai.,
al. , 1996).
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 361

Small magnetostriction is important not only for static magnetic softness


but also for reducing the eddy current loss at high frequency ranges. Magnetic
core properties under various frequencies and inductions were examined for
Fego Zr,
the nanocrystalline Fe90 FeB9 Hh B, and Fes,
Zr7 B3 , Fesg Nb,7B
FeB' Nb Bg
9 alloys subjected to
anneal ing at 923 K for 3.
3.66 ks (Suzuk
(Suzukii et aI., 1993). Core loss (W) under
various measurement conditions, saturation magnetic induction (B s)' relative
(/-Ie)' coercivity (He), electrical resistivity (p), sample
permeability (IJe)'
t) and density (Om) for the nanocrystalline Fe-M-B alloys are
thickness ((t)
summarized in Table 9. 2. The data for amorphous Fe,s Fe7B Sig
Si 9 B 13 (METGLAS
2605S2) annealed at 643 K for 3.6 ks are also listed for comparison. The W
values attained for the nanocrystalline alloys are smaller by about 10% - 50% 50 %
than those for the amorphous Fe-Si-B alloy. Given the fact that core losses
generally increase with either increasing coercivity or decreasing electrical
resistivity, the results shown in Table 9.2, i. e. , higher He and lower p, for
the nanocrystalline alloys compared with amorphous Fe-Si-B, seem to be
inconsistent with the lower core losses in the nanocrystalline alloys. The
reason for this inconsistency lies in the small As of the nanocrystalline alloys.

Table 9.2 Magnetic properties (W, B s ' lJe' He and As)' electrical resistivity (p), ribbon
thickness ( Fe-M-B (M = Zr, Hf and Nb) and
t) and density (Om) for nanocrystalline Fe-M-8
amorphous Fe-Si-8 alloys (Suzuki et al. ,1993).
Fe-Si-B
Fego Zr,
Fe90 83
Zr7 B FeSg Hh
FeS9 Hf, B,
8, g
Nb, 8 9
Fes, Nb7B Fe7S Si
Fe,s Sig9B
8 13
,3
Structure
Nanocrystalline Nanocrystalline Nanocrystalline Amorphous
WW50
w\l)50(W/kg)
(W/kg) 0.21 0.14 0.19 0.24
W\J),oo (W/kg) 0.82 0.61 0.97 1.22
W\J),k (W/kg)
W\J)" 2.27 1.70
170 2.50 3.72
W~)lOOk
WmOOk (W/kg) 79.7 59.0 75.7 168
B,CT) 1.63
163 1. 59
1.59 1.49 1.56
1. 56
lJe at 11 kHz 22,000 32,000 22,000 9,000
He (A/m)
(Aim) 5.6 5.6
56 8.0 2.4
A,(( xX 10 6 )
A, -1.
-1.11 -1.2
-12 0.1 27
p( xX 108 Om) 44 48 58 137
t (llm)
(~m) 18 17 22 20
Om( x 3
X 10- kg/m )) 33
7.62
762 8.46 7.74
774 77. 18
(1) Walp
(1) WalP represents core loss at (XX
a: x 1O- T and (3f3 Hz.
10- 1 T

The eddy current loss of core materials are composed of classical (We)
( W a ) portions. The classical eddy current loss is calculated
and anormalous (W
from
("IT tfB m ) 2
W = (IT (9.39)
e 6pO m

The ratio of the classical eddy current loss to the total eddy current loss is
often defined as the anomaly factor (T)
(T/) :
362 Kiyonori Suzuki

(9.40)

Figures 9. 12 and 9. 13 show the changes in the loss components at B m = = 1. 0 T


as a function of frequency for the nanocrystall ine Fego Zr7 B3 and the amorphous
Fe78 Si gB 13 alloys, respectively. Although We for the amorphous Fe-Si-B alloy
Fe78SigB13
with a higher p is smaller than that for the nanocrystalline Fe-Zr-B alloy, W a of
the former alloy is extremely high in high frequency ranges, leading to the
lower values of the total core loss in the nanocrystalline Fe-Zr-B alloy. The 17 T)
value at 50 kHz is 5. 7 and 1. 4 for the amorphous Fe-Si-B and the
nanocrystalline Fe-Zr-B alloys, respectively. The small 17 -- 1 for the
T) '"
nanocrystalline Fe-Zr-B alloy is comparable to that (1.5) of the commercial
zero-magnetostrictive Co-based amorphous alloy. Inomata et al. (1983)
reported that the anomaly factor in the frequency range 10 - 50 kHz for
amorphous Fe-Si-B-Nb alloys is proportional to the magnetostriction. In their
study, the 17 T) value of the amorphous Fe-Si-B-Nb alloys increased linearly from
7. 5 with increasing As from 7 x 10- 6 to 30 X 10- 6 (Fig. 9.14). The small
1.5 to 7.5
anomaly factor of the nanocrystall ine samples arising from the small As readily
nanocrystalline readi Iy
explains their excellent high frequency core characteristics.
25
,~
bJ)
OJ.)
Zr 78 -'
Fe90Zr7B3
Feyo
B m=1.0T
"'-'"-,"
~ 20
E
E Anomalous e.c.loss
~
.,
pg.
~C-
:::: 15 Classical e.c.loss
~ai
Uu
....
>.
u
,.) Static hysteresis loss
....
~
]0
10
"'
Q)
a.
e-
If.
:2
en
.2
.!2
~ 5
0
u
0
]0 1
10 10 2 103 10'
Frequency J (Hz)
.f
Figure 9. 12 Changes in separated core losses at a maximum induction of 1.0 T in core
loss per cycle (W pe ) for nanocrys1alline
nanocrystalline FegoZr,B 3 as a function of frequency (f) (Suzuki
et ai.,
al., 1993).

Magnetic cores are often embedded in an epoxy resin in practical use.


Hence, the influence of residual stress, caused by molding the samples in an
epoxy resin, on the magnetic softness is of technological importance. This
effect was examined for nanocrystalline (Fel-xCox)90Zr7B3 with a range of As
values (Suzuki et al., 1994a). The residual stress in this experiment was
estimated to be - 1. 4 XX 10 8 Pa from the stress induced magnetic anisotropy
(0.67kJ/m 3 at Ass =3.1 x 10- 6 )).. Figure 9.15 shows the changes in Ile at
1 kHz and 1 MHz for the samples before and after molding as a function of As
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 363

25
~
Fe78Si9BI3
FenSi9B13
CJJ
bJl
..:.: Bm=1.0 T
""---...,...., 20 Anomalous e.c.loss
5-5
~ 15
i5. Classical e.c.loss\
\
\
U ">,
;>.,
Static hysteresis loss \
\ .
.......u<lJ
U
<l) 10 \
0-
0..

'"enen
2'"
..s
1: 5
~
0
u
0
10 1 ]022
10 ]033
10 10'
Frequency,f
Frequency,! (Hz)

Figure 9. 13 Changes in separated core losses at a maximum induction of 1. 0 T in core


loss per cycle ( W pc) for amorphous Fe7s
Fe?8 Si g B
8 13
'3 as a function of frequency (f) (Suzuki
et al. , 1993).

10
f~IO
f=10 to 50 kllz
kHz
= 0.3 T
Bm=
8

"">:-..:..:
U
u
0
,9
6
~
>,
;>.,

""'ES
0
<:: 4
~'"

2
,
/
;/

,,
//
/
/
;/

0 10 20 30 40
6)
As (X 10-6)
As(XIO

Figure 9. 14 Change in the anomaly factor (1]) (1)) in amorphous Fe-Si-8-Nb


Fe-Si-B-Nb alloys as a
function of saturation magnetostriction o.s) al. , 1983).
Us) (Inomata et al.,

The smallest influence of the residual stress is confirmed at i\Ass = 0 and a high
IJee value above 10 4 is obtained even after molding in an epoxy resin. The small
IJ
magnetostriction is one of the greatest advantages of Fe-based nanocrystalline
alloys.
364 Kiyonori Suzuki

105.
,- - - - - - - - - - - - - - - - - - - - - - ,

" "kHz
~
.:,~ore At
(Before ~
molding)
moldi'g)
~
0
... 0 a
-~~
II'
I
~ 0

: \.
\- At I kHz
I \ ' (After molding)
At 1
I MHz

(Before m~Olding)
i~\
,.~
:"'"
t., :"'"
I \

~~ ", ~.o
.0
o 01 I 0 "........
"...... -
- ---_
__...._
-

-
I
I f
I .

/.\
I
/-\

I
.......__
-___
---~_
--~_
0

.
-

... .
,.
/I
AI //
/

..,/
___",
. ' ./.
..JI/
/. At I MHz (After molding)

I02';-
J02';-- --';;-
';;- --+- -----
~---~--~ ~
-4 -2 0 2 4
As (XIO-
.Ie, (x 10-6 )

Figure 9. 15 Relation between relative permeability (fJe) (lJe) and saturation magnetostriction
0.,) (Fe, - xCox) 90 Zr, B3 ( X = 0 to 0.05)
CAs) for nanocrystalline (Fe'-xCox)90Zr7B3(X=0 O. 05) alloys before and after resin
molding (Suzuki et ai.,
al. , 1994a).

9.4 Principles Underlying Alloy Design

9. 4. 1
9.4.1 Alloying Elements and Alloy Systems

In this section, we first overview various nanocrystalline soft magnetic alloys


reported to date. Second, the principles underlying alloy design in these
nanocrystalline
nanocrystall ine soft magnetic alloys are discussed. Some of the reported
alloys whose magnetic core properties are similar or inferior to those of
Fe-based amorphous alloys (i. e. , lJe at 1 kHz < 10 4 or He> 10 Aim) are left
out in this discussion.
Major alloy systems reported so far may be listed chronologically as:
Fe-Si-B-Nb-Cu (Yoshizawa et ai., aI., 1988), Fe-Si-B-Nb-Au (Kataoka et al. ,
1989a), Fe-Si-B-V-Cu (Sawa and Takahashi, 1990), Fe-(Zr or HO-B (Suzuki
etal., 1990), Fe-(Ti, Zr, Hf, Nb or Ta)-B-Cu (Suzuki et al., aI., 1991b),
Fe-Si-B-(Nb, Ta, Mo or W)-Cu (Yoshizawa and Yamauchi, 1991), Fe-P-C-
(Mo or Ge)-Cu (Fujii et ai.,
al., 1991), Fe-Ge-B-Nb-Cu, Fe-Si-B-(AI, P, Ga or
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 365

Ge)-Nb-Cu (Yoshizawa et al., aI., 1992), Fe-AI-Si-Nb-B (Sendust-Nb-B)


(Watanabe et a!.,aI., 1993), Fe-AI-Si-Ni-Zr-B (Supersendust-Zr-B) (Chou
al., 1993) and Fe-Si-B-Nb-Ga (Tomida, 1994).
et a!., 1994) . More recently,
a!., 1995)
nanostructural formation was reported for Fe-Si-B-U-Cu (Sovak et al.,
al., 1996) alloys, however, the coercivity
and Fe-Si-B-Nd-Cu (Muller et a!.,
AI m.
values of these alloys were 10 to 20 Aim.
Table 9.3 summarizes the alloy composition, mean grain size, saturation
magnetization, coercivity and effective permeability
permeabil ity of the developed
materials. Most of these alloys can be grouped empirically into the following
formula:

[Fe
FM EM
, eu]ML LM

~~] [:1 ~i
~~I
Co No J66-91 [Ti V

: [
Zr Nb P (Ag)
Hf Ta W 2-8 Ga Ge 2-31 Au 0-1
(9.41)
where the elements given in bold letters are found most commonly in the
developed alloys. FM in Eq. (9. 41) are ferromagnetic elements. The
approach to improve magnetic core properties by nano-crystallization is
insignificant for Co- or Ni-rich systems from a technological viewpoint since
small K 1 and As values have already been realized in Co-rich amorphous or Ni-
rich Permalloy. Hence, Co and Ni were used exclusively as additives in Fe-
based nanocrystalline systems. EM denotes early transition metals, i. e. IVa
to Via metals, which are the slowest diffusive elements in the system (Frank
al. , 1994), providing retarded crystal growth (Koster et al.,
et a!., a!., 1991). ML
are metalloid, semi-metal and simple metal elements. LM are late transition
metals whose enthalpy of mixing with Fe is positive. The developed alloys
listed in Table 9. 3 may be divided into Fe-metal based FM-EM-ML and Fe-
metalloid based FM-EM-ML-LM systems in terms of alloying elements. Here,
the principles underlying alloy design in these 2 types of nanocrystalline soft
magnetic alloys are discussed respectively.

Table 9. 3 Mean grain size (D),


CD), saturation magnetization (M,),
CM s ), coercivity C He) and
(He)
O. 4 Aim (lie)
relative permeability at 1 kHz under 0.4 C11.) of various nanocrystalline soft magnetic
materials.

Composition D(nm)
DCnm) MsCT)
MsCn He (Aim)
HeCA/m) lie ( 10 3 )
I1.(10 References

FeSS-91 ZrS_7
Zr5-7 B
8 22-6
_6 12 to 18 1.6
.6 to 1.7 4 to 8 10 to 31 (Suzuki
CSuzuki et al. ,
1991a, 1996)
Fes7-91 s_78
FeS7-91 Hf5 B2-7
2- 7 13 to 14 1.65 4 to 6
.55 to 165 10 to 32 (Suzuki
CSuzuki et al. ,
1991a, 1993)
Feso-ss
FeSO.S5 Nb5_7B
s.78 S._14 9 to 10 1.4
.4 to 1.55 6 to 8 10 to 38 (Suzuki
CSuzuki et al. ,
1993)
366 Kiyonori Suzuki

Table 9.3 Continued from page 365.

Composition O(nm) Ms(T)


MsCT) Hc(A/m)
Hc(A!m) /.1.(10 3)
lJe(10 References

Fess. 7-SS.S
7-S9. sCo
Coo.s
o.s_1.3 Zr7 B3
83 15 1.65
65 to 1.7 4 to 5 23 to 27 (Suzuki et al. ,
1994a)
Fess-ss Zr7 B
8 3Ab_s 10 to 15 1.5 to 1 6 10 to 17 (Inoue, 1996)
Fess-ss Zr7 B
8 3Sb_s 10 to 15 1.5 to 1.6 10 to 14 (Inoue, 1996)
FeS3 Nb7B
Fes3 89
sGa, 1.48 4.
4.88 38 (Makino et al.
1995)
FeS3 Nb7B
Fes3 89
sGe, 1.47
1. 47 5.6
56 29 (Makino et al. ,
1995)
Fe70 Ah
AI, SilO Ni 4Zr2 B
8s 20 5 26 (Chou et al.
1993)
Fe66 AisSi'4 Nb3Bs
Nb389 14 0.75 1.2 30 (Watanabe et al. ,
1993)
Fes9_74
Fe69_74 Si
Si,o_'3
lO _'3 B
89
sNb3Ga4-6 10 to 20 1.1
1. 1 to 1.3 1.2 (Tomida
CTomida,, 1994)
Fes2
FeS2 Tb B
8 10
IO Cu, -10 1.39 to 1
1.55
55 3.5 to 4.9
4. 9 11 to 20 (Suzuki et al. ,
1991b)
Fes2-s,
FeS2_9' Zrs_s B'_'2
8'-'2 Cu, 10 to 12 1.25
1. 25 to 1.65
1. 65 2.4 to 8 10 to 48 (Suzuki et al. ,
1991b)
Feso Hf78
Fe90 Hf7B2 Cu,
2Cu, 10 1.6 3.0 18 (Suzuki et al. ,
1991b)
FeS4 Nb7B
Fes4 8 sCu, -10 1.48
.48 to 1
1.56
56 8.6 to 8.9 13 to 16 (Suzuki et al. ,
1991b)
1991 b)
Fes2
FeS2 Ta7BlOCu,
Ta7 8 lO Cu, -10 .46
46 8.9
89 11 (Suzuki et al. ,
1991b)
Fes6 8 6Cu,
FeS6 Zr4Nb3B .54 3.7 18 (Suzuki et al. ,
1991a)
Fes4
FeS4 Zr35
Zr3.sNb3.5
Nb35 B8 sCu, 8 1.53
53 1.7 100 (Makino et al. ,
1995)
FeS6 Zr4 Nb3B
Fess 86
sAu, .51
51 6. 11 (Suzuki et al. ,
1991a)
FeS6 Zr4Nb 3B
Fes6 8 6Pd, .54 5.
5 10 (Suzuki et al. ,
1991a)
FeS6 Zr4 Nb3B
Fes6Zr4 8 6Pt, 1.47
1. 47 7.7 10 (Suzuki et al. ,
1991a)
Fe72 Si
Fen Si'3 8 sV6Cu,
13 BsV6Cu, 23 4 15 (Sawa and
Takahashi, 1990)
Fe73.5 Si
Fe73.s 13 589
Si'3.s Nb3Cu,
BsNb3CUI 12 1.24
1. 24 0.5 100 to (Yoshizawa et al. ,
150 1988)
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 367

Table 9.3 Continued from page 365.

Composition O(nm)
D(nm) M,(T)
MsCT) He (Aim) /.1.(10
IJ 3
3
e ( 10 )) References

Fe73.5 Si '55sB
FensSi,s B,Nb
7Nb3Cu,
3Cu, 1. 15 05 150 (Yoshizawa and
Yamauchi, 1989)
B'3 Nb3Cu,
Fes, Sb B13
Fe8' 1.55 12.8
128 9 (Yoshizawa and
Yamauchi, 1989)
Fe73.5Si,uB9 Ta3Cu,
FensSi13sB9 11 40 (Yoshizawa and
Yamauchi, 1991)
Fe73.5 Si
Fens 13 . 5B
Si'3S B99M03CUI
3Cu, 15 70 (Yoshizawa and
Yamauchi, 1991)
Fe73.5 Situ
Fens Si'3.s B 9W 3CUI
B9W3Cu, 15 70 (Yoshizawa
CYoshizawa and
Yamauchi, 1991)
Fe73.5 Situ
Fens Si'3.s B 9Nb3Au,
B9Nb3Au, - 3 20 (Kataoka et al.
1989a)
Fe,s
Fe78 PIG
p'G C22 Si
Sio CUO.5s 16
Ge3 CUo
sGe3
o. 5 1.8 (Fujii et al. ,
1991)

9.4.2 Alloy Design in Fe-Metal Based Nanocrystalline Alloys


In earlier investigations of the Fe-metal based systems, bee-Fe precipitates with
grain sizes smaller than L~x were obtained upon primary crystallization of ML-free
Fe60 C030 ZrlO (Guo et al. , 199])
amorphous Feso 1991) or Fen Zrs (Suzuki et ai.,
aI., 1994b).
However, their soft magnetic properties (iJe"'" 103) were inferior to those of Fe-
(lJe '" 10
based amorphous alloys, making these ML-free alloys technologically
insignificant. The low iJelJe of the nanocrystalline Fen
Fe92 Zrs is due to a large volume
fraction of residual amorphous phase with a small exchange stiffness constant
(large V am and small A am). The poor magnetic softness, given the fine grain size
in this B-free nanocrystalline sample, is readily understood by considering the
effects of these two factors on (K 1 > in Eq. (9. 30b).
< K 1) 30 b). Although the volume
fraction of the residual amorphous phase can be reduced by increasing annealing
temperature, the secondary crystallization process commences at relatively low
annealing temperatures in the B-free Fen Zrs alloy, leading to the formation of
magnetically hard Fe3 Zr. Hence, a wide temperature interval between primary
and secondary crystallization is the key to realizing a magnetically soft
nanostructure. It is worth remarking here that nanoscale microstructures can be
produced even in FM-EM binary systems.
Obvious conditions for the occurrence of wide temperature intervals between
primary and secondary crystallization in amorphous precursors are as follows:
The redistribution of glass forming elements between the primary precipitates and
the amorphous matrix is prosperous upon primary crystallization; these glass
forming elements can induce a dramatic increase in the thermal stability of the
368 Kiyonori Suzuki

matr ix amorphous phase. A study reported by Ohnuma et al. (1981)


matrix (198 1) showed
that the crystallization temperature of an amorphous Fe-Zr-B ternary system
increases dramatically with increasing B content and the maximum crystallization
temperature exceeds 1000 K. Moreover, the solubility limit of B in <x-Fe ex-Fe is
virtually zero, and, so, B is enriched remarkably in the residual amorphous phase
(Zhang et ai.,aI., 1996). Hence, the thermal stability of the residual amorphous
phase in primary crystallized FM-EM systems is enhanced dramatically by the
addition of a small amount of B. For example, the temperature interval of the
primary and secondary crystallization processes in amorphous Fe91 Zr7 B2 is as large
as 240 K, significantly greater than that of the B-free amorphous Fe92ZrS 100 K).
Fe92 Zrs ((100 K) .
Since B is one of the most effective elements at increasing the Curie
temperature in Fe-rich amorphous alloys (0' Handley, 1983), another
significant effect induced by the addition of B is the strengthening of the
intergranular magnetic coupling
coupl ing (i. e., enhancement of A am) am ) through an
n
increase in n m . In fact, the soft magnetic properties of amorphous Fe-Zr
alloys after primary crystallization are improved dramatically by the addition of
B, and nanocrystalline Fe91Zr7B2 exhibits a high IJe J.1e value of 31 ,000. Thus, the
three reasons which have made B the most commonly used additive among all
the ML elements are CD the effectiveness of B in enhancing the thermal stability
of the residual amorphous phase, (ll '1! the effectiveness of B in increasing the
Curie temperature of the residual amorphous phase and @ the small solubility
of B in <x-Fe.
ex-Fe.
As we find in the random anisotropy model, <K 1) 1> may vary as Ki K1 and,
hence, a rational approach to better soft magnetic properties is to reduce the
intrinsic magnetocrystalline anisotropy in the bee-Fe
bcc-Fe nanocrystallites. Hayashi
et al. (1987) reviewed various thin-film crystalline materials and summarized
that K 1 = a0 is expected in Fe-Ga-Si, Fe-AI-Ge, Fe-Co-Ga-Si and Fe-Co-AI-Ge
I = Fe-Co-Al-Ge
Fe-Al-Si. Therefore, the addition of AI, Si,
besides classical Fe-Ni and Fe-AI-Si.
Ga and Ge to FM-EM-B may result in a further reduction of < 1>. Examples of
< K 1).
this approach can be found in nanocrystalline Fe66 Ai s Si 14 Nb 3B9 ,
Fe7oAbSilONi4Zr2Bs and Fe69-74Fe69.74 Si ,o- Nb3Ga4_6 and their minimum He
,3 B9 Nb3Ga4.6
10.13
<1.2 Aim) is truly smaller than that of FM-EM-B alloys (4 Aim). However, as
(1.2
we discussed in Section 9. 3. 3, dynamic soft magnetic properties are often
governed by the magnetostriction constant rather than < K I) 1 > when < K I) 1 > is
suppressed to below a certain level. Accordingly, the IJe J.1e values at 1 kHz in
Table 9.3 for the Fe-(Zr, Hf or Nb)-BNb) -B and Fe-AI-Si-Nb-B
Fe-Al-Si-Nb-B alloys are similar. In
addition, the increase in ML content results in a serious decrease in the
saturation magnetization, mak ing some of the developed alloys featureless
making
from the viewpoint of technological applications.

9.4.3 Alloy Design in Fe-Metalloid Based Nanocrystalline Alloys


The first nanocrystalline soft magnetic alloy reported by Yoshizawa et al.
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 369

(1988), Fe735Si,35B9Nb3Cul
FensSi13sB9Nb3Cul or commercially known as FINEMET, belongs to
this type, and the research on FM-EM-ML-LM type nanocrystalline alloys has
the longest history in the field. Yoshizawa and Yamauchi (1990) showed that
Cu-free amorphous Fe745 135 B9Nb3 crystallizes into a mixture of Fe23 B6 and
Fe74S Si 13s
Fe3 B phases with a grain size of about 100 nm after annealing anneal ing at 823 K. The
precipitates after annealing change from these Fe-B compounds to nanoscale
bcc-Fe(Si) by the addition of 1 at. %Cu. Noh et a!. al. (1990) showed that the
onset temperature of the crystallization of bcc-Fe(Si) is reduced preferentially
by the addition of Cu in this system and explained the microstructural change
induced by Cu. The key aspect in the FM-EM-ML-LM alloys is the role played
by the LM element.
As we discussed in the previous section, a magnetically soft nanostructure
can be prepared even in ML-free Fe-( Zr, Hf or Nb) -B ternary systems.
Nb )-B
However, the addition of Cu becomes essential in these alloy systems when
the B content is relatively high. The soft magnetic properties in the Fe-Zr-B
system deteriorate markedly with increasing B content above 8 at. % to
10 at. %, because the increase in B content above this level results in the
formation of magnetically harder compounds upon primary crystallization
al. , 1991 a). However, this formation of compounds is suppressed
(Suzuki et a!.
by an addition of 1 at. % Cu and the compounds-free magnetically soft
nanostructure is restored. As a result, the compositional range for the
excellent magnetic softness expands significantly towards the B-rich region.
Given the fact that the B content of the Fe-Si-B-Nb-Cu alloy studied by
Yoshizawa et a!., al., and Noh et a!. al. (1990), was 9 at. %, the observed
microstructural changes induced by Cu in the Fe-Si-B-Nb-Cu and the high boron
content Fe-Zr-B alloys seem to be due to the same origin. This implies
possibilities of developing Cu-free FINEMET by reducing the metalloid content.
However, a larger amount of EM elements would be required to maintain
sufficient glass forming ability in such metalloid reduced compositions for
FINEMET. Consequently, this approach may end up with the compositions of
NANOPERM.
Hono et al. (1992)
( 1992) studied the microalloying effect of Cu on the primary
crystallization process in the Fe-Si-B-Nb-Cu (Hono et a!. al. , 1992) and Fe-Zr-B-
Cu (Zhang et a!., al., 1996) alloys by means of APFIM. They detected Cu-
enriched clusters in the precursor amorphous phase at early stages of
annealing (far before the crystallization event took place). These Cu clusters
are presumed to act at the heterogeneous nucleation sites for primary bee-Fe bcc-Fe
precipitates. The cluster formation of Cu was considered to be due to the
positive enthalpy of mixing (b.H (llH mix ) between Fe and Cu and hence, similar
effect on the nanocrystallization process in Fe-Si-B-EM based alloys is
expected for other LM elements whose b.H llH mix with Fe is also positive. This
approach is exemplified in nanocrystalline Fen5 Fens Si 13s
135 B9 Nb 3Au, (Kataoka
et ai.,
a!. , 1989a). Although the b.H llH mix between Ag and Fe is positive, Ag could
not be added to Fe-rich melt-spun alloys because of its strong immiscibility in
370 Kiyonori Suzuki

Fe even in the Iiquid


liquid state. Hence, Cu remains the most technologically
significant additive among the LM elements.
The role played by Cu in the FM-EM-ML-LM alloys can be understood by
taking into account the effect of Cu on crystallization kinetics. Kataoka et al.
((1989b)
1989b) reported that the apparent activation energy of crystall ization for
amorphous Fe74s gNb 3 decreases from 4.6 to 3.9 eV by an addition of
13 . 5B9
Fe745 Si 13s
1 at. % Cu. The heterogeneous nucleation sites induced by the Cu cluster
formation seem to be the mechanism of the reduced activation energy. Since
this reduction of the activation energy is accompanied by the preferential
precipitation of bee-Fe, copper appears to accelerate preferentially the
nucleation kinetics of bee-Fe. The onset temperature of crystallization in Fe-
metalloid amorphous phase tends to decrease with increasing the average
outer electron concentration (e / a) of Fe in the amorphous phase (Donald et
al. , 1982). Hence, elements with high outer electron concentrations, such as
Pd and Pt, may also accelerate the crystallization kinetics of primary bee-Fe.

9. 5 Prospects

Nanocrystalline soft magnetic materials prepared by crystallization of


amorphous precursors can mostly be described as FM 66 - 91 EM 2 - 8 ML 2 - 31 LM o- 1 ,
FM66-91EM2-sML2-31LMo-l,
where FM are ferromagnetic elements, EM denotes early transition metals,
ML includes metalloid, semi-metal and simple metal elements and LM are late
transition metals. The highest saturation magnetization (M s ) of 1.7 T in the
nanocrystalline soft magnetic alloy family was found in Feg! Fe91 Zr7 B2 2 with effective

permeability (J..Ie) '" 30, 000. On the other hand, the highest value of lJ.e
(JJ.e) -- J..Ie--
'"
150 ,000 was obtained for Fe735Si!35BgNb3Cul
Fe73sSi13sB9Nb3Cul with M s of 1.25 T.
A key parameter for good magnetic softness, besides small grain size,
small K 11 and large exchange stiffness, is volume fraction of the residual
amorphous phase (V am)' Small V am is preferable when the Curie temperature
(n
of the residual amorphous phase (nm ) is substantially lower than that of the
nanocrystall
nanocrystalline ( T'f{) , whereas moderate V am may reduce < K 1
ine phase (Tn, n
1>> if n m is
close to Tg.
T'f{.
The coercivity of nanocrystalline soft magnetic alloys near T '" -- n m could
vary as the 6th power of the spontaneous magnetization of the intergranular
region. This suggests that further improvements of the soft magnetic properties
are possible in alloy systems with relatively low n mm if the Curie temperature of
the residual amorphous phase is enhanced. Elements that preferentially
partition between the nanocrystalline bee-Fe and the residual amorphous
phases following nanostructural formation may have great effect upon the
intergranular magnetic couplcoupling.
ing.
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 371

References
Alben, R.,
R. , J.
J.J.
J. Becker and M.C.
M. C. Chi. J. Appl. Phys. 49: 1653 (1978)
Bozarth, R. M. Ferromagnetism. D. Van Nostrand Co. , Inc. , Princeton, p.
Bozorth,
831 (1951)
Chikazumi, S. Physics of Magnetism. John Wiley & .Sons, Inc., Inc. , New York,
1964
Chou, T. M. Igarashi and Y. Narumiya. J. Magn. Soc. Jpn. 17: 197 (1993)
Donald, I. W. H. A. Davies and T. Kemeny. J. Non-Cryst. Sol Solids
ids 50: 351
(1982)
Frank, W.,
W. , A. Horner, P. Scharwaechter and H. KronmOlier.
Kronmuller. Mater. Sci.
Eng. A 179/180: 36 (1994)
Fujii, Y., H. Fujita, A. Seki and T. Tomida. J. Appl. Phys. 70: 6241
( 1991)
Gleiter, G. Prog. Mater. Sci. 33: 223 (1989)
Guo, H. Q. , T. Reininger, H. KronmOller,
Kronmuller, M. Rapp and V. Kh. Skumrev.
Sol idi A 127: 519 (1991)
Phys. Status Solidi
Hasegawa, N. and M. Saito. J. Magn. Soc. Jpn. 14: 313 (1990)
Hasegawa, N. , M. Saito, N. Kataoka and H. Fujimori. J. Mater. Eng. Per.
2: 181 (1993a)
Hasegawa, N. Doctoral Thesis. Tohoku University, Sendai, unpublished, p.
216 (1993b)
Hayashi, H., M. Hayakawa, W. Ishikawa, Y. Ochiai, H. Matsuda, Y.
Iwasaki and K. Aso. J. Appl. Phys. 61: 3514 (1987)
Hernando, A.A.,, M. Vazquez, T. KulikandC.
Kulik and C. Prados. Phys. Rev. B 851:
51: 3581
(1995))
(1995
Hernando, A., P. Marin, M. Vazquez, J. M. Barandiaran and G. Herzer.
Phys. Rev. B 858:
58: 336 (1998)
Herzer, G. IEEE Trans. Magn. MAG-25: 3327 (1989)
Herzer, G. IEEE Trans. Magn. MAG-26: 1397 (1990a)
Herzer, G. Vacuumschmelze GmbH Internal Research Report 071/70.
Vacuumschmelze GmbH, Hanau, unpublished (1990b)
Herzer, G. Scr. Metall. Mater. 33: 1741 (1995)
Herzer, G. J. Magn. Magn. Mater. 157/158: 133 (1996)
Herzer, G. In: K. H. J. Buschow ed. Handbook of Magnetic Materials.
Elsevier Science, Amsterdam, Vol. 10, Ch. 3, p. 415. (1997)
Ho, K. Y., X. Y. Xiong, J. Zhi and L. Z. Cheng. J. Appl. Phys. 74: 6788
((1993)
1993)
Hoffmann, H. J. Appl. Phys. 35: 1790 (1964)
Hoffmann, H. IEEE Trans. Magn. 9: 17 (1973)
Hoffmann, H. and T. Fujii. J. Magn. Magn. Mater. 128: 395 (1993)
Hono, K., K. Hiraga, Q. Wang, A. Inoue and T. Sakurai. Acta Metall.
Mater. 40: 2137 (1992)
372 Kiyonori Suzuki

Inomata, K., M. Hasegawa, T. Kobayashi and T. Sawa. J. Appl. Phys.


54: 6553 (1983)
Inoue, A. Mater. Sci. Forum 225 - 227: 639 (1996)
225-227:
Ishiwata, N. , C. Wakabayashi and H. Urai. J. Appl. Phys. 69: 5616 (1991)
Ito, S., K. Aso, Y. Makino and S. Uedaira. Appl. Phys. Lett. 37: 665
(1980
(1980))
Kataoka, N. , T. Matsunaga, A. Inoue and T. Masumoto. Mater. Trans. JIM
30: 947 (1989a)
Kataoka, N. , A. Inoue, T. Masumoto, Y. Yoshizawa and K. Yamauchi. Jpn.
J. Appl. Phys. 28: Ll820
L1820 (1989b)
Kim, S.H.,
S. H. , M. Matsuura, M. Sakurai and K. Suzuki. Jpn. J. Appl. Phys.
32: 676 (1993)
Koster, U. and U. Herold. In: H. J. Guntherodt and H. Beck. Glassy Metals
I. Topics in Applied Physics. Springer Verlag, Berlin, p. 225 (1981)
Koster, U. , U. Schunemann, M. Blank-Bewersdorff, S. Brauer, M. Sutton
and G.B. Stephenson. Mater. Sci. Eng. A 133: 611 (1991)
Lee, E. W. Rep. Prog. Phys. 18: 184 (1955)
Loffler, J. F. , H. B. Braun and W. Wagner. J. Appl. Phys. 85: 5187 (1999)
Luborsky, F. E. In: F. E. Luborsky, ed. Amorphous Metallic Alloys.
Butterworth & Co Ltd. , London, p. 360 (1983)
Makino, A., A. Inoue and T. Masumoto. Mater. Trans. JIM 36: 924 (1995)
Muller, M. and N. Mattern. J. Magn. Magn. Mater. 136: 79 (1994)
Muller, M., N. Mattern and U. Kuhn. J. Magn. Magn. Mater. 157/158: 209
( 1996)
Noh, T. H. , M. B. Lee, H. J. Kim and I. K. Kang. J. Appl. Phys. 67: 5568
(1990)
O'Handley, R.C. Solid State Commun. 21: 1119 (1977)
0' Handley, R. C. In: F. E. Luborsky, ed. Amorphous Metallic Alloys.
Butterworth & Co Ltd. , London, p. 257 (1983)
Ohnuma, S., M. Nose, K. Shirakawa and T. Masumoto. Sci. Rep. RITU
A 29: 254 (1981)
Sawa, T. and Y. Takahashi. J. Appl. Phys. 67: 5565 (1990)
Schafer, R. , A. Hubert and G. Herzer. J. Appl. Phys. 69: 5325 (1991)
Shimizu, 0., K. Nakanishi and S. Yoshida. J. Appl. Phys. 70: 6244 (1991)
Slawska-Waniewska, A., M. Gutowski and H. K. Lachowicz. Phys. Rev.
B 46: 14,549 (1992)
Sovak, P., P. Petrovic, P. Kollar, M. Zatroch and M. Knoc. J. Magn.
Magn. Mater. 140-144: 427 (1995)
Stoner, E. C. and E. P. Wohlfarth. Phil. Trans. Roy. Soc. A 240: 599
(1948)
Suzuki, K., N. Kataoka, A. Inoue, A. Makino and T. Masumoto. Mater.
Trans. JIM 31: 743 (1990)
Suzuki, K.,
K. , A. Makino, N. Kataoka, A. Inoue and T. Masumoto. Mater.
Trans. JIM 32: 93 (1991a)
Processing and Modeling of Novel Nanocrystalline Soft Magnetic Materials 373

Suzuki, K., A. Makino, A. Inoue and T. Masumoto. J. Appl. App!. Phys. 70:
6232 (1991b)
Suzuki, K., M. Kikuchi, A. Makino, A. Inoue and T. Masumoto. Mater.
Trans. JIM 32: 961 (1991c)
Suzuki, K., A. Makino, A. Inoue and T. Masumoto. J. Appl. App!. Phys. 74:
3316 (1993)
Suzuki, K. , A. Makino, A. Inoue and T. Masumoto. J. Magn. Soc. Jpn. 18:
800 (1994a)
K. , A. Makino, A. P. Tsai, A. Inoue and T. Masumoto. Mater. Sci.
Suzuki, K.,
Eng. AI79/180: 501 (1994b)
Suzuki, K.,
K. , J. M. Cadogan, V. Sahajwalla, A. Inoue and T. Masumoto. J.
Appl. Phys. 79: 5149 (1996)
App!.
Suzuki, K.,
K. , G. Herzer and J. M. Cadogan. J. Magn. Magn. Mater. 177-
181: 949 (1998a)
Suzuki, K. and J. M. Cadogan. Phys. Rev. B58: 2730 (1998b)
Taneko, N. , Y. Shimada, K. Fukamichi and C. Miyakawa. Jpn. J. Appl.
L 195 (1991)
Phys. 30: L195
Tomida, T. Mater. Sci. Eng. AI79/180: 521 (1994)
D. G. and E.P
Tonge, D.G. E. P Wohlfarth. Phil. Mag. 3: 536 (1958)
Varga, L. K. , L. Novak and F. Mazaleyrat. J. Magn. Magn. Mater. 210:
L25 (2000)
Watanabe, H. , H. Saito and M. Takahashi. J. Magn. Soc. Jpn. 17: 191
(1993)
Yoshizawa, Y., S. Oguma and K. Yamauchi. J. Appl. Phys. 64: 6044
((1988)
1988)
Yoshizawa, Y. and K. Yamauchi. J. Magn. Soc. Jpn. 13: 231 (1989)
Yoshizawa, Y. and K. Yamauchi. Mater. Trans. JIM 31: 307 (1990)
Yoshizawa, Y. and K. Yamauchi. Mater. Sci. Eng. A133: 176 (1991)
Yoshizawa, Y. Y. Bizen, K. Yamauchi and H. Sugihara. Trans. lEE Jpn.
112A: 553 (1992)
Yoshizawa, Y. Doctoral Thesis. Tohoku University, Sendai, unpublished, p.
160(1993)
Zhang, Y. , K. Hono, A. Inoue and T. Sakurai. Scr. Mater. 34: 1705 (1996)

The author is grateful to the Australian Research Council for its continuing financial support,
and is indebted to his collaborators who have contributed to the results mentioned in this
chapter. They include Associate Professor N. Kataoka, Professor A. Makino, Professor A.
Inoue, Professor T. Masumoto, Associate Professor J. M. Cadogan and Dr. G. Herzer.
Index

ac susceptibil ity 139, 140, 143, 146, characteristic length scales 67, 123,254
147,213 circular nano-dots 127, 130
activation volume 35, 36, 43 - 45, 82, coercivity 1, 13, 15, 16, 19 - 26, 35-
1,13,15,16,19 35 -
103,110,133-135,314 39,43,45,46,52,148,172,174,175,178,
amorphous 25,26, 38 - 40,
25,26,38 88 , 204 , 216
40,88,204,216 188 - 191 ,193
, 193 - 195, 199,202 - 207 ,211 ,
-220,223,224,228,230,231,234,237, 215 - 226,228,229,232
226 , 228,229,232 - 236,238,240,
264,270, 275, 323, 324, 327, 328, 332 -
264,270,275,323,324,327,328,332 241 , 249, 253, 254, 265, 271 , 278, 281 ,
334,338 - 340 ,342 - 344 ,346 ,347 ,349- 310,321 , 323, 325, 326, 331 , 334, 336 -
354 340,343,346,347,352
angular dependence 16,77 - 79 ,,82
82 , 84 , collective spin 105, 111 - 113
105,111-113
87,89,90,92,93,99, 103, 104, 111 , 112,
87,89,90,92,93,99,103,104,111,112, Constricted-Wall 33
301,311 Core loss 343,344
anisotropy 1,3,9 - 12,14 - 18,22 - 27, 27 , Coulomb blockade 75
33,35,36,38 - 40,42,44,45,52,64,67, crossover temperature 108 - 111 , 161 -
68,73,77 ,79 - 82,84 - 86,88 - 90,94 - 165
96, 105 - 109,123,127,130,133,137,
96,105 109, 123, 127, 130, 133, 137, crystal structure 15, 184, 185, 187, 195,
138,140-144,149,151,161-163,165,
138,140 - 144,149,151,161- 163,165, 197,198,208,216,223,240,325
172,174- 176, 178 - 181,187 - 189,193
176,178 crystalline electric field 253
- 195,201,202,204,207,208,212,215-
-195,201,202,204,207,208,212,215- crystallization 209,216,218,223,224,
217, 226, 232, 234, 240, 241 , 249, 253, 226, 228, 231 , 236, 271 , 275, 323, 324,
254,256,257,267,273,274,278,298 - 327,332,340,347,349 - 352
300,306,321,323,325,326,328,330 - Curie behavior 95
332,334 - 338 ,340,345,350 Curie temperature 1, 2, 6 - 8, 24, 27,
anisotropy energy 10,16,40,73,77
10, 16,40 ,73 ,77 ,80, 32,46,60,62,64,178,199,202,203,209
81 ,84 - 86,
86 , 91 , 100, 121 , 129, 144, 161 , - 213,216,223,224,226,253,338 - 340 ,
235,267, 274, 297, 323, 325, 326, 328 -
235,267,274,297,323,325,326,328 350,352
332,334 - 336,338
336 ,338 curling mode 21,22, 36,
36,91
91 - 93, 113,
anisotropy energy barrier 105,137 179
anisotropy field 15,16,20,21,25,78, distribution of switching fields 96, 97,
108,112,120,121,128,129,145,148,
108, 112, 120, 121, 128, 129, 145, 148, 299
151 , 162 - 164, 177,179,187,206,207,
151,162 177, 179, 187, 206, 207 , domain wall 5, 16 - 18, 20, 22,
22 , 31 , 33,
33 ,
209,286,296 34,40,54,67,72,73,75-77 ,93,94, 103,
,93,94,103,
average grain size 39, 180, 204, 211 , 105, 109, 110, 114, 123 - 125, 189, 300, 300 ,
213,225,236 310,325,326,329,331
306 - 310 ,325,326 ,329 ,331
avoided level crossing 106, 107 domain wall width 73,93,94,105,123,
73,93,94, 105, 123,
Brillouin function 95 300,326,331,337
Index 375

domains 5, 16 - 18,20,35,38
18, 20 , 35 , 38 - 40,67,
40, 67 , 191-193,204,212,215,217-220,223-
72,73,116,121,123,124,126,130,133, 227, 232, 233, 235, 239, 268, 271 , 273,
227,232,233,235,239,268,271,273,
134, 173, 181 ,229,305,317,325
134,173,181,229,305,317,325 280,282, 323, 325, 328, 330 - 337, 346,
effective anisotropy 21 , 36, 77 , 80, 81 , 347,349,350,352
108,133,215,330,331 Hall magnetometry 284,307
electron holography 75,285,293 Hall probe magnetometry 75
energy barrier 18, 23, 27,
27 , 40 - 44, 77 , hard magnetic phase 172, 178, 189,
81 ,82,95 - 99, 104, 105, 108, 110, 111 , 217,228,235,236,239,271
119,133-135,138-141,143-148,151,
119,133 - 135, 138 - 141 , 143 - 148,151 , hysteresis loop 3,11 ,13,
, 13, 14, 19,24,26,
162, 163,306,307,312 - 314
155 - 159, 162,163,306,307,312 34,38,39,72-
34,38,39,72 - 74,76,78,79,82,93,94,
energy level spectrum 106
106 103,104,106,107,148,153,164,173,
exchange 2 - 9, 11 , 12, 14, 15, 17, 17 , 18, 174,178,180,192,194,195,203,204,
22,29,31,32,34,35,39,46,60,65,73, 225,230,235,236,238,271,273,335
77 ,91,103,112,120,121,123,125-127, intrinsic coercivity 179, 189, 211 , 218,
171 - 174, 178, 180 - 184,189,190,200,
171-174,178,180 184, 189, 190,200, 219,223 - 225 ,227 ,228 ,233
203 - 207,211
207 , 211 - 215,217,219,220,223, Josephson junction 76,285
224,226,228 - 230 ,234 - 236 ,239 - 241,
236,239 241 , Langevin function 95, 148,150,153
148, 150,153
253,254, 256, 263 - 265, 267, 268, 271 , Lorentz microscopy 75,285,306
273 - 276,297,300,327,328,332 - 334, macroscopic quantum tunnel ing 77 , 105,
339,340,349,352 111,160,161,164-167
exchange correlation length 325, 328 - magnetic domain 75,76,229
330,333 - 335,337,338
335 ,337,338 magnetic force microscopy 32, 72, 75,
exchange coupling 32,39 , 63, 163, 172,
32,39,63,163, 109,113,279,283,292
177 , 179 - 181,183,188,189,193,204,
177,179 181 , 183, 188, 189, 193, 204, magnetic linear dichroism 75
205,212 - 217,219,220,223,224,226 - magnetic molecular clusters 105
229,232,233,235,236,238
229,232, 233, 235, 236, 238 - 240,253, 240, 253, magnetic nanoparticles 103, 142.
142, 161,
254,273,274,276 166,280,306
exchange interaction 4, 6, 19, 37, 60, magnetic nanowires 62,133
67,80,108,111,125,212,215,224,228, magnetic properties 1, 13, 34, 40, 48,
229, 235, 254, 256, 267, 275, 276, 300, 54,59,63,64,66,75,103,133,184,187-
301,329,335,338 189,193,200,202,203,207,211,216 -
exchange length 18,19,22,73,91
18 , 19 ,22 ,73 ,91 , 105, 226,228 - 230,232 - 234,236,239,253,
123,300,304,331 254, 257, 269, 276, 279, 280, 282, 291,
254,257, 291 ,
exchange spring behavior 189 323,334,340,343,349 - 352
exponential prefactor 99 magnetic relaxation 82, 106, 137, 147,
82,106,137,147,
Extrinsic Properties 1,2, 13, 14,35,46
1 ,2,13, 149,154-156,158,161,164-166
Fe-Si-B-Nb-Cu 324,325, 327,328,346,
324,325,327,328,346, magnetic viscosity 41 , 43, 156 - 158,
351 164,165,312-314
Fe-Zr-B 324,327,335,344,349,351 magnetism 1 - 6,8,9,14,16,18,29,
Finite Element Discretization 124 30 ,37 .39,41 ,44,47
30,37,39,41 ,44 47 - 49,52,54,60,62,
49,52 , 54 , 60 , 62 ,
FMR 162,273 - 275 63,65,69,72,98,114,142,166,167,171,
Fokker-Planck equation 41 ,95
41,95 172,180,181,208,235,249,279,315,
172,180,181, 208, 235, 249, 279, 315,
giant spin model 106 318,324,352
Gilbert equation 95, 119, 120, 122,235, magnetization 1,3 - 7, 9 - 11 , 13 - 17,17 ,
266,307 20 - 22,24 - 27,32,35,
27,32, 35,38 38 - 42,44.46,
42,44, 46,
grain size 19,32,
19, 32, 171, 172, 181, 189, 47,60,62,67,68,72- 87,91 - 93,95-
376 Index

100,103- 114, 116, 119, 121 - 125,128-


100,103-114,116,119,121-125,128- 293,297,314
130,132 - 135, 137,138,142,144 - 146, nanostructure 1, 17, 25, 29 - 32, 35, 37
148-153,155,156,158-161,163-166,
148- 153,155,156,158-161,163- 166, 40 , 46, 51 , 54,58,60,63,65,75,
- 40,46,51, 54, 58,60,63, 65, 75, 114,
115, 119, 137, 138, 164, 171 - 174, 177-
115,119,137,138,164,171 177 -
171-178,181-
171- 178, 181 - 183,188,189,191,193,
179,184,188,190,193,202,204,206,
200,202, 204, 206, 207, 210 - 214, 216,
212,220, 222, 226, 232, 234, 239 - 241 ,
218 - 220, 222, 228, 230, 232, 235, 236, 247,250,264,280,301,319,323,325,
247,250, 264, 280, 301, 319, 323, 325,
240,241 , 253, 254, 256 - 258, 260, 261 , 334,336,338,349,351
265 - 268 ,270 - 276 ,279 ,285 ,286 ,288- Neel-Brown model 95 - 99, 102, 103,
290,292 - 308 , 31 0 - 315
308,310 ,323 - 328
315,323 ,331 ,
328,331 110,111
337,338,340,346,347,350,352 ohmic damping 109
optimum grain size 228
magnetization dynamics 27 ,40
27,40
parity effect 107
magnetization reversal 22, 36,72
36, 72 - 77 , permanent magnet 1, 2, 10, 12, 14, 16,
82 - 84 ,87,91
,87 ,91 - 99 , 102 - 104,
104 , 109 - 111
111,, 19,20,23 - 25 ,29,30,32,34,38
,29,30,32,34 ,38 - 40,48,
40 ,48 ,
113, 114, 119, 126, 127, 134, 191, 194, 49,72, 114,171 - 174,177,178,184,188,
49,72,114,171 174, 177 , 178,184, 188,
279,281 ,285,293 - 296 ,298 - 315
279,281,285,293 189, 195,200 - 203,206 - 208 , 216,217,
magnetocrystalline anisotropy 1, 3, 4, 234,240,243,249,250,253,254,257,293
11 , 12, 17,24,
11,12, 17, 24 , 36,46,64,77
36, 46 , 64, 77 , 80,
80 , 82, 85,
permeability
permeabil ity 25, 172, 265, 271, 323,
324,343,345 - 347 , 352
86,93,130,133,187-189,200,217,228,
86,93,130,133,187- 189,200,217,228,
phase diagram 60 , 184 - 186, 195, 196,
60,
296,323,325 - 332 332,334
,334 - 340,350 198,208,247,250,257
magnetoresistance 25,54 ,65 ,75
25,54,65,75 quantum tunneling 27,73,74,81, 96,
27,73,74,81,96,
magnetron sputtering 206, 211,
21 1, 234, 104,105,107,109,110,114,116,137,
235,264,270 138 , 144 , 159 , 161 , 162 , 164 - 166,306
166 ,306
maximum energy product 178,188, 193, random-anisotropy 29, 32, 33, 39, 171 ,
172,175,180
203 - 207 ,,209,219,220,222
209,219,220,222 - 228 ,232-
rapid quenching 190, 216, 219, 220,
234,241,254,264
227,233
mechanical alloying 30, 32, 172, 190, recording media 10 , 25, 26, 31 , 38, 41 ,
10,
203,205,206,211,223,225,226,232 137,149,160,188
melt-spinning 172 ,203,2111,220,264
172,203,21 ,220,264 remanence enhancement 29, 32, 39,
"mesoscopic" physics 104 171 , 172, 180, 189, 195, 205, 215 - 217,
micromagnetic simulation 91,92,
91, 92, 119, 224,230,235,238 - 240
shape anisotropy 11 ,24,36,43,77 , 79,
11,24,36,43,77
123,126,130,301
80,82,85,86,88 - 90,92, 133, 188,296,
molecular magnets 114 311 ,314 ,325
31 1,314
multilayer film 207,235,236,238,264 -
207,235,236,238,264- single-domain particles 27, 74, 76, 91 ,
266,276 153 ,280,3111
113, 153,280,31
nanocomposite magnet 184, 190, 211
211,, small particles 15-17,20,21,31,35,
15- 17,20,21,31,35,
215,218 - 221 , 225, 226, 228
228,, 229, 232
232,, 39 ,41 ,44,86,91 , 113,
39,41 113 , 127,279,284,286,
127,279 ,284 ,286,
295,298,301,304,306
295,298,301 ,304,306 - 308
233,238,240,253,254,264,265,269,276
soft magnetic phase 189,202,215,226,
nanocrystalline 40, 79, 93, 109, 172,
228,229,234,236,275
212,217, 222, 224, 228, 230, 323 - 325, spin operator 106
327,328,330,332,334 - 347 ,349 - 352 spin reorientation 212, 229, 253, 257,
nanoparticle 27, 30, 31 , 72, 75, 76, 81 , 258
82 ,88,89,91 ,95,98,99,
82,88,89,91 ,95,98 , 99, 105, 107 - 111,
111 , spin-dependent tunneling 75
113,114,141,142,143,146,147,151,
113, 114, 141, 142, 143, 146, 147, 151, SQUID 72,74 -76,
- 76, 78, 83, 88, 90, 92,
153,158,163,165,166,279,281,292, 93,98,
93 ,98, 102 - 104 , 109,
109 , 113, 118 , 139,285,
139,285 ,
Index 377

302,303 telegraph noise measurements 96 - 99


STM-assisted CVD 296,299 ,314
296 ,299,314 telegraph noise spectroscopy 110
stochastic fluctuations 98 thermal activation 19,22,41 ,77 ,95,96,
Stoner-Wohlfarth model 16, 20, 77 , 79, 109-
109-111,111, 114, 119, 128, 138, 144, 149,
80,82,90,95,300,307,311,326 158,279, 300, 301 , 305 - 309, 311 , 313,
superparamagnetism 44,95,139 314
surface anisotropies 10,31 ,33 thin film 17,29 - 31 ,51 ,54,55,62,63,
switching field 45,75,77 , 79,81 ,82,84
79,81,82,84 67,93,94,98,119,123,125 - 127,130, 127, 130,
- 94 ,96 - 100, 102 - 104, 109, 110, 112, 187, 188, 204, 234 - 237, 240, 270, 281 ,
130, 172, 178, 204, 294, 295, 298 - 301 , 302,336
304,305,307,308,311,313,314 thin film elements 119, 127, 130
127,130
switching field measurements 82, 86, tunnel splitting
spl itting 107
96,97,103,104,109 uniform rotation mode 77,82,86,88,92,
77 ,82,86,88,92,
switching probability 96, 98, 99, 102, 279
110 vibrating reed magnetometry 75
switching rate 97,108,110 Waiting Time Measurements 96,98,99
Handbook of Advanced
Magnetic Materials
Volume II: Advanced Magnetic Materials:
Characterization and Simulation
Handbook of Advanced
Magnetic Materials
Volume II: Advanced Magnetic Materials:
Characterization and Simulation

Edited by:

Yi Liu
Center for Materials Research and Analysis
University of Nebraska
Lincoln, Nebraska

David J. Sellmyer
Center for Materials Research and Analysis
University of Nebraska
Lincoln, Nebraska

Daisuke Shindo
Institute of Multidisciplinary Research for Advanced Materials
Tohoku University
Sendai, Japan

@ Tsinghua University Press ~ Springer


Library of Congress Cataloging-in-Publication Data

ISBN-IO: 1-4020-7983-4 e-ISBN-IO: 1-4020-7984-2


ISBN-13: 978-1402-07983-2 e-ISBN-13: 978-1402-07984-9

2006 Springer Science+Business Media, Inc.


All rights reserved. This work may not be translated or copied in whole or in part without the
written permission of the publisher (Springer Science+Business Media, Inc., 233 Spring Street, New
York, NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis.
Use in connection with any form of information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter developed is
forbidden.
The use in this publication of trade names, trademarks, service marks and similar terms, even if they
are not identified as such, is not to be taken as an expression of opinion as to whether or not they are
subject to proprietary rights.

Printed in the United States of America.

9 8 76 54 3 2 I SPIN 11097730

springeronline.com
Handbook of Advanced Magnetic Materials
Preface

In December 2002, the world's first commercial magnetic levitation supertrain


went into operation in Shanghai. The train is held just above the rails by mag-
netic levitation (maglev) and can travel at a speed of 400 km/hr completing
the 30km journey from the city to the airport in minutes. Now consumers are
enjoying 50 GB hard drives compared to 0.5 GB hard drives ten years ago. A-
chievements in magnetic materials research have made dreams of a few dec-
ades ago reality. The objective of this book is to provide a comprehensive re-
view of recent progress in magnetic materials research. The whole book con-
sists of four volumes, each volume focusing on a specific field. Graduate
students and professional researchers are targeted as the readers. Each chap-
ter will have an introduction to give a clear definition of basic and important
concepts of the topic. The details of the topic are then elucidated theoretically
and experimentally. New ideas for further advancement are then discussed.
Sufficient references are also included for those who wish to read the original
work. Many of the authors are well known senior scientists. We have also
chosen some accomplished young scientists to provide reviews on new and ac-
tive topics.
In the last decade, one of the most significant thrust areas of materials re-
search has been nanostructured magnetic materials. There are several critical
sizes that control the behavior of a magnetic material. For example, the coer-
civity of a magnetic material made of particles increases with decreasing parti-
cle size, reaching a maximum where coherent rotation of a single-domain par-
ticle is realized, and then decreases with further decrease of the particle size.
For a composite made of a magnetically hard phase and soft phase, when the
grain size of the soft phase is sufficiently large, the soft and hard phases re-
verse independently. However, when the grain size of the soft phase is re-
duced to a size of about twice the domain wall thickness of the hard phase, the
VI Preface

soft and hard phases will be exchange-coupled and behave as if a single mag-
netic phase is present. Such behavior can be used to increase the energy
product of high-performance permanent magnets. Size effects become critical
when dimensions approach a few nanometers, where quantum phenomena ap-
pear. The first volume of the book has therefore been devoted to the recent
development of nanostructured magnetic materials, emphasizing size effects.
Our understanding of magnetism has advanced with the establishment of
the theory of atomic magnetic moments and itinerant magnetism. In general,
the magnetism of a bulk material can be considered as the superposition of a-
tomic magnetic moments plus itinerant magnetism due to conduction electrons.
In practical applications the situation becomes much more complicated. The
boundary conditions have to be taken into account. This includes the size of
the crystals, second-phase effects and intrinsic properties of each phase. The
effects of magnetic relaxation over long periods of time can be critical to un-
derstanding. Simulation is a powerful tool for exploration and explanation of
properties of various magnetic materials. Simulation also provides insight for
further development of new materials. Naturally, before any simulation can be
started, a model must be constructed. This requires that the material be well
characterized. Therefore the second volume of the book provides a compre-
hensive review of both experimental methods and simulation techniques for the
characterization of magnetic materials. After an introduction, each section gives a
detailed description of the method and the following sections provide examples
and results of the method. Finally further development of the method will be
discussed.
The success of each type of magnetic material depends on its properties
and cost which are directly related to its fabrication process. Processing of a
material can be critical for development of artificial materials such as multilay-
er films, clusters, etc. Moreover, cost-effective processing usually deter-
mines whether a material can be commercialized. In recent years processing
of materials has continuously evolved from improvement of traditional methods
to more sophisticated and novel methods. The objective of the third volume of
the book is to provide a comprehensive review of recent developments in pro-
cessing of advanced magnetic materials. Each chapter will have an introduc-
tion and a section to provide a detailed description of the processing method.
The following sections give detailed descriptions of the processing, properties
and applications of the relevant materials. Finally the potential and limitation
of the processing method will be discussed.
The properties of a magnetic material can be characterized by intrinsic
properties such as anisotropy, saturation magnetization and extrinsic proper-
Preface vn
ties such as coercivity. The properties of a magnetic material can be affected
by its chemical composition and processing route. With the continuous search
for new materials and invention of new processing routes, magnetic properties
of materials cover a wide spectrum of soft magnetic materials, hard magnetic
materials, recording materials, sensor materials and others. The objective of
the fourth volume of this book is to provide a comprehensive review of recent
development of various magnetic materials and their applications. Each chap-
ter will have an introduction of the materials and the principals of their applica-
tions. The following sections give a detailed description of the processing,
properties and applications. Finally the potential and limitation of the materials
will be discussed.
NASA is considering the launching of spacecraft by maglev. The first
stage rocket, which accounts for two-thirds of the cost and is lost every
launch, would be replaced by a maglev track. Using a 50 ft track NASA scien-
tists have accelerated a model spacecraft to 96kph in less than half a second.
In the last few decades the knowledge of mankind has been expanding rapidly
into deep space measured by light years and the nano world where building
blocks of atoms are being engineered. Magnetism and magnetic materials are
among the most intriguing and fascinating science and engineering fields. Un-
doubtedly advances in magnetic materials research will continue to fuel our un-
derstanding of the universe in the new century. We hope this book will provide
a useful reference for researchers working at the frontier of magnetic materials
research.
We would like to express our sincere thanks to all our devoted authors,
technical editors, and publishers for making this book possible.

The editors
Contents

Preface V
List of Contributors XIX

1 Advanced Magnetic Force Microscopy Tips for Imaging Domains 1


1. 1 Introduction ... ... ... ... ... ... 1
1.2 Magnetic Force Microscopy 2
1. 2. 1 Basic Operating Principles of MFM 2
1.2.2 Force Gradient Detection 2
1.2. 3 Sensitivity and Resolution 3
1.2.4 Tip-Sample Interaction 4
1.2.5 Force Sensor 7
1.3 Development of Advanced MFM Tips 8
1. 3. 1 High Coercivity CoPt MFM Tips 8
1.3.2 Superparamagnetic and Low Stray Magnetic Field
MFM Tips 13
1. 3.3 Electron-Beam-Induced-Deposited
(EBID)MFM Tips 15
1. 3.4 Point-Dipole MFM Tips 17
1.3.5 Focused Ion Beam Milling MFM Tips 20
References 22

2 Lorentz Microscopy and Holography Characterization of Magnetic


Materials 24
2. 1 Introduction'" 24
2.2 Instrumentation 25
2. 3 Analytical Electron Microscopy for Structure
Characterization 29
2. 3. 1 Outline of Current Analytical Electron
Microscopy'" 29
2. 3.2 Thickness Measurement by EELS 31
2. 3. 3 Elemental Mapping with EDS 33
2.4 Lorentz Microscopy on Magnetic Domain Structure 35
X Contents

2. 4. 1 Principles of Lorentz Microscopy ... ... ... ... ... ... ... ... ... 35
2.4.2 Lorentz Microscopy Using Conventional Transmission
Electron Microscopes '" 36
2.4.3 Lorentz Microscopy Using Scanning Transmission
Electron Microscopes ..... , ... ... ... ... ... ... ... 44
2.5 Principles and Application of Electron Holography'" 46
2. 5. 1 Principles of Electron Holography 46
2. 5.2 Practice of Electron Holography 51
2. 5. 3 Appl ication of Electron Holography 52
2. 6 Concluding Remarks '" 63
References 63

3 Characterization of Magnetic Materials by Means of Neutron


Scattering '" 66
3. 1 Introduction'" 66
3. 1. 1 Cross Section Formalism 69
3. 1.2 Polarized Neutron Beam Instrumentation 71
3. 1.3 The Polarization of the Scattered Beam 77
3. 2 Elastic Magnetic Scattering ... 80
3.2. 1 Small-Angle Scattering 80
3.2.2 Neutron Diffraction 84
3.2.3 Reflection of Neutrons from Magnetic Surfaces and
Interfaces 90
3. 3 Inelastic Magnetic Scattering 97
3. 3. 1 Studies of Elementary Excitations by Triple-axis
Spectroscopy .. .. .. .... .. .. .. .. .. .. .. .. .. .. .. .. .. 97
3.3.2 Detection of Slow Motions by Neutron Spin Echo 101
3. 4 Summaries'" 107
References ... ... ... 108

4 Advanced Transmission Electron Microscopy of Nanostructured


Magnetic Materials 113
4. 1 Introduction 113
4.2 Specimen Preparation 114
4.2.1 Bulk Samples 114
Contents Xl

4.2.2 Magnetic Thin Films 114


4. 2. 3 Magnetic Nanowires '" 118
4. 2.4 Magnetic Powders 119
4.2.5 Special Techniques 120
4.3 Electron Diffraction 121
4. 3. 1 SAD Pattern and the Ring Pattern '" 121
4. 3.2 Convergent Beam Electron Diffraction 122
4.3.3 Nanodiffraction 124
4.4 High Resolution and Super-Resolution TEM .. , 129
4.4. 1 An Image Processing Model ... ... ... 130
4. 4.2 Procedure for Image Reconstruction 135
4.4.3 Test of the Image-Processing Model 135
4. 5 Selected Reflection Imaging ... ... ... 138
4.5. 1 Origination of the Technique 138
4. 5.2 Experimental Method in Conventional TEM 139
4.5.3 Application of SRI .. .. .. .. 140
4. 5.4 Experimental Set-Up in a STEM 142
4.6 STEM and Z -Contrast Imaging '" 143
4.7 Electron Energy Loss Spectroscopy 145
4.8 Concluding Remarks 148
References ... ... ... ... ... ... 148

5 Mossbauer Spectroscopy Characterization of Soft Magnetic


Nanocrystalline Alloys 151
5. 1 Introduction 151
5.2 Mossbauer Spectroscopy 155
5.2. 1 Principles of the Mossbauer Effect 155
5.2.2 Hyperfine Interactions 164
5.2.3 The rf-Mossbauer Technique 168
5. 3 Experiment ... ...... ... ... ... ... ... ... ... 171
5.3.1 Formation of the Crystalline Nanostructure in
Amorphous Matrix ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 171
5. 3.2 Primary Characterization of Thermal Stability of Amorphous
Precursor-Formation of Nanostructure 173
5. 4 Mossbauer Study of the Structure and Magnetism of FINEMET
XII Contents

Alloys . 178
5. 5 M6ssbauer Study of the Structure and Magnetism of
NANOPERM Alloys .. 188
5. 5. 1 Conventional M6ssbauer Studies .. 191
5. 5. 2 The rf-M6ssbauer Studies . 212
5.6 Surface and Bulk Nanocrystallization of Amorphous FeCuNbSiB
and Fe-M-B-Cu Alloys-the CEMS Study . 232
5. 7 Short Range Order in Amorphous Precursors-the
rf-M6ssbauer Study '" . 244
5. 8 Determination of the Grain Size by XRD and SAXS
Techniques-Correlation with the M6ssbauer Results . 250
References . 260

6 Atom Probe Characterization of Microstructures of Nanocrystalline


and Nanocomposite Magnetic Materials . 266
6. 1 Introduction . 266
6.2 The Atom Probe Technique . 267
6. 3 Microstructural Evolution in Nanocrystall ine Soft Magnetic
Materials . 271
6.3.1 FINEMET CFe-Si-B-Nb-Cu) . 271
6.3.2 NANOPERM CFe-Zr-BC-Cu . 279
6.3.3 HITPERM CCFe,Co)-Zr-B-Cu) . 287
6.4 Microstructural Evolution in Nanocomposite Magnets . 289
6. 4. 1 Fe3 B/Nd2 Fe'4 B System . 290
6.4.2 a-Fe/Nd2 Fe14 B System . 295
6.4.3 Amorphous Remaining a-Fe/Nd2 Fe14 B
Nanocomposite . 299
6.5 Roles of Microalloyed Elements in Nanocrystallization . 301
6.6 Effect of Heating Rates on Nanocrystalline Microstructure
Evolution . 304
6. 7 Summary . 306
References . 306

7 Itinerant-Electron Metamagnetism 310


7. 1 Introduction .. ... ... ... ... ... ... 310
Contents XII

7. 2 Theoretical Aspects of Itinerant-Electron Metamagnetism 311


7. 2. 1 Landau Expansion Coefficients and Magnetic Phase
Diagram 311
7.2.2 Paramagnetic Susceptibility Maximum in the
Temperature Dependence 314
7. 3 Itinerant-Electron Metamagnetism of Laves-Phase
Exchange-Enhanced Pauli Paramagnets 315
7.3. 1 Metamagnetic Transition in the Ground State 315
7.3.2 Relationship Between the Susceptibility Maximum
and the Transition Field 317
7.3.3 Metamagnetic Transition at Finite Temperatures 320
7.4 Correlation Between the Magnetovolume Effects
and Metamagnetic Transition ... ... ... ... ... ... ... ... ... ... ... ... ... ... 321
7.4. 1 Concentration Dependence of the Curie Temperature
and Spontaneous Magnetization .. .... .. .. .... .... 321
7. 4.2 Pressure Effects on the Curie Temperature and
Spontaneous Magnetization ... ... ... ... ... ... ... ... ... 322
7. 4. 3 Thermal Expansion Anomaly and Spin Fluctuations 327
7.5 Determination of the Landau Coefficients 329
7.5. 1 Pressure Effect on the Critical Field of the
Metamagnetic Transition ... ... ... ... ... ...... ... ... ... ...... 329
7.5.2 Comparison Between the Experimental and Theoretical
Magnetic Phase Diagrams 333
7. 6 Suppression of Spin Fluctuations in Laves-phase
Metamagnets 337
7.6. 1 Concentration Dependence of the Specific Heat Coefficient
of Laves-Phase Compounds ...... ...... ...... ...... ... ... ... 337
7.6.2 Large Electronic Specific Heat Coefficient Due to Spin
Fluctuations and Its Suppression Under High Fields 338
7.7 Metamagnetic Transition at Finite Temperatures of Ferromagnetic
LaCFe'-xSix)13 NaZn13-type Compounds 340
7.7. 1 Magnetization and Magnetic Phase Diagram 340
7.7.2 Thermal Expansion Anomaly 346
7.7.3 Pressure Effect on the Metamagnetic Transition 347
7.7.4 Control of the Metamagnetic Transition by Hydrogen
XIV Contents

Absorption , '" .,. 350


7.8 Drastic Changes in Magnetic and Electrical Properties and
Their Practical Applications 353
7.8. 1 Isotropic Giant Volume Magnetostriction 353
7.8.2 Giant Magnetocaloric Effect 358
7. 8. 3 Giant Magnetoresistance 365
7. 9 Concluding Remarks 366
References : 367

8 Modeling of Hysteresis in Magnetic Materials 372


8. 1 Introduction '" , 372
8.2 Development of Model Theories of Hysteresis 373
8.3 Magnetism at the Discrete Level of Individual Atoms and Beyond
to the Continuum Level: Landau-Lifschitz-Gilbert Model and
Micromagnetics 375
8.4 Magnetism of Domain Rotation: Stoner-Wohlfarth 378
8. 5 Magnetism at the Level of Domain Boundaries:
Neel. Globus-Guyot. Bertotti 382
8.6 Magnetism at the Macroscopic Scale: the Integration of Single
Domain Switching Processes and the Preisach Model 386
8. 7 Magnetism at the Multidomain Level: Energy Considerations
and the Jiles-Atherton Model 391
8.7. 1 Description of the Anhysteretic Magnetization 392
8. 7.2 Extension to Describe Hysteresis 393
8. 7. 3 Extension to Describe the Effects of Stress
on Magnetization 395
8.7.4 Extension to Describe the Effects of Frequency on
Magnetization 396
8.7.5 Applications'" 397
8. 8 Summary 403
References 404

9 Coarse-graining and Hierarchical Simulation of Magnetic Materials:


the Fast Multipole Method 407
9. 1 Introduction ... ... ... 407
Contents XV

9.2 The Fast Multipole Method: Simplest Implementation 408


9. 3 Cartesian Formulation of the FMM 414
9.4 History of the FMM 416
9. 5 Micromagnetic Applications of the FMM 417
References 420

10 Numerical Simulation of Quasistatic and Dynamic Remagnetization


Processes with Special Applications to Thin Films and
Nanoparticles 421
10. 1 Basic Micromagnetic Concepts and Main Energy
Contributions 421
10.2 Discretization Methods: Simplicity and Speed
Versus Exact Shape Approximation 422
10.2. 1 Regular CTranslationally Invariant) Grids 423
10.2.2 Tetrahedron Mesh 424
10.3 Evaluation of Various Energy Contributions 427
10.3. 1 Anisotropy Energy in Polycrystalline Samples 428
10.3.2 Exchange Energy: Node-supported Discretization,
Heisenberg-Like Form and Angle-based
Interpolation ... ... ... ... ... ... ... ... ... ... ... 429
10.3.3 Stray Field Evaluation on Regular Grids 434
10.4 Energy Minimization Methods 443
10.4.1 Standard Minimization Technique: Conjugate
Gradients ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 444
10.4.2 Equation of Motion Techniques and Simple
Relaxation Methods 447
10.4.3 Advanced Relaxation Methods Combined with the
Extrapolation Techniques 449
10. 4.4 AI ignment Methods ... ... 451
10.5 Equilibrium Magnetization Structures and Quasistatic
Remagnetization Processes 454
10.5. 1 Nanosized Magnetic Elements 454
10.5.2 Extended Thin Films and Patterned Structures 459
10.5.3 Quasistatic Remagnetization in Nanocomposites:
Individual Particle Switching and Cooperative
XVI Contents

Remagnetization Processes 463


10.6 Equilibrium and Non-Equilibrium Thermodynamics: Langevin
Dynamics, Monte Carlo Method and Path Integrals 469
10.6. 1 Fast Remagnetization Processes: Langevin
Dynamics vs. Monte Carlo Method 469
10.6.2 Slow Remagnetization Dynamics 489
References 503

11 Preisach Model and Simulation of Relaxation Kinetics 508


11 . 1 Introduction ... ... ... ... ... 508
11 .2 The Response Operator 509
11. 3 The Preisach Model ... ... 510
11. 4 Ensembles of Systems in Random Potential 518
11.5 Representability of the Ensemble Evolution by the PM 520
11.6 Connection of the "Classical" Irreversibility Parameters
with the PM ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 523
11 .7 Representations of Some Ensembles by the PM 525
11.7.1 Campbell Random Potential and Stabilization of
Domain Walls ... ... ... ... ... ... ... ... ... ... 525
11 . 7.2 Periodic Potentials with Random Phase 530
11.8 Uncertainty in Prediction of Relaxation Kinetics
Based on the PM '" 535
11 . 9 Summary , 539
References 540

12 Antiferromagnetism of Mn Alloys 541


12. 1 Introduction 541
12.2 Theory of Itinerant Electron Magnetism with the
Hubbard Model 542
12. 2. 1 Model for the Itinerant Electron Systems 542
12.2.2 Path-Integral Approach for the Itinerant-Electron
Magnetism 542
12.2.3 Saddle Point (Molecular Field) Approximation 544
12.2.4 Rotation of the Local Spin Axes in the Complex
Magnetic Structures 545
Contents xvn

12.2.5 Magnetic Excitation Energy and the Exchange


Constant 547
12.3 First Principles Approach for the Magnetic Structures of
Transition Metal Systems 550
12.3. 1 Tight-binding (TB) -LMTO Method for Complex
Magnetic Structures 550
12. 3. 2 Coherent Potential Approximation (CPA) for
Disordered Alloys 552
12. 3. 3 Effective Exchange Constant 553
12.4 Electronic and Magnetic Structures of y-Mn and
Mn alloys ... ... .. . 554
12.4. 1 y-Mn 554
12.4.2 FeMn Disordered Alloy .. .. .. 559
12. 4. 3 L 10 -Type MnPt, MnNi and MnPd Alloys 564
12.4.4 L1 2 -Type and y-Phase Mn31r Alloys 572
12.5 Experimental Observations of Antiferromagnetism
of Mn Alloys ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 578
12. 5. 1 Concentration Dependence of the Neel Temperature
of L 12 -type (=y' -Phase) Mn Ordered
Alloy Systems ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 578
12. 5. 2 Concentration Dependence of the Neel Temperature
of y-phase Mn Disordered Alloy Systems ...... ...... 581
12.5.3 Lattice Distortions and Spin Structures of
y-phase Mn Disordered Alloys 582
12.5.4 Concentration Dependence of the Neel Temperature
of Ll o (=CuAu-l)-type Alloy Systems 586
12.5.5 Low-temperature Specific Heat and Temperature
Dependence of Electrical Resistivity ...... ... ... ...... 587
12. 6 Antiferromagnetic Properties Related to Magnetic
Devices 589
12.6. 1 Magnetovolume Effects and Thermal Strains 590
12.6.2 Blocking Temperature and Magnetic Domains in
Antiferromanets 593
12.6.3 Frustration of Antiferromagnetic Spin Structures
XII Contents

and Exchange Bias Field 594


12. 7 Concluding Remarks 597
References 597

Index 602
List of Contributors

1. Sy-Hwang Liou Department of Physics and Astronomy and


Center for Material Research and Analysis
University of Nebraska
Lincoln. Nebraska 68588-0111
sliou@unlnotes01. un1. edu
2. Daisuke Shindo Institute of Multidisciplinary Research for Advanced
Young-Gil Park Materials,
Tohoku University, Sendai 980-8577, Japan
shindo@iamp. tohoku. ac. jp
3. G. Ehlers Institute Laue-Langevin, 6, rue Jules Horowitz, BP
156
38042 Grenoble Cedex 9, France
F. Klose ehlers@ill. fr
Oak Ridge National Laboratory, Spallation Neutron
Source Project
9700 South Cass Avenue, Argonne, IL 60439, USA
fklose@anl. gov
4. Yi Liu Center for Materials Research and Analysis
J. Sellmyer University of Nebraska,
Lincoln, NE 68588-0113, USA
Yliu@unlserve. un!. edu
5. Michat Kopcewicz Institute of electronic materials technology,
01-919 warsaw, poland
Kopcew-m@sp. itme. edu. pi
6. Kazuhiro Hono National Institute for Materials Science
1-2-1 Sengen, Tsukuba 305-0047, Japan
Hono@nrim. go. jp
7. Kazuaki Fukamichi Department of Materials Science,
Graduate School of Engineering,
Tohoku University,
Aoba-yama 02, Sendai 980-8579, Japan
fukamich@material. tohoku. ac. jp
XX List of Contributors

8. D. C. Jiles Ames Laboratory,


X. Fang US Department of Energy
W. Zhang and Department of Materials Science and Engineering,
Iowa State University, Ames, IA 50011
gauss@ imap. ameslab. gov
xwfang@ee. ucla. edu
9. P. B. Visscher Department of Physics and Astronomy and
Center for Materials for Information Technology
University of Alabama, Tuscaloosa AL 35487-0324
pv@pi. ph. ua. edu
10. D. V. Berkov INNOVENT e. V. , Jena, Germany
N. L. Gorn db@ innovent-jena. de
11. K. L. Metlov Institute of Physics ASCR, Na Siovance 2, Prague
8, CZ-18221
metlov@fzu. cz
12. Akimasa Sakuma Electronic & Magnetic Materials Research Lab. ,
Hitachi Metals Ltd. , Japan
Akimasa_Sakuma@hitachi-metals.co. jp
Kazuaki Fukamichi Department of Materials Science, Graduate School
of Engineering, Tohoku University, Sendai, Japan
fukamich@material. tohoku. ac. jp
1 Advanced Magnetic Force Microscopy Tips for
Imaging Domains

Sy-Hwang Liou

1. 1 Introduction

Magnetic force microscopy (MFM) is a tool for imaging domains and studying
a variety of local magnetic phenomena. It has been widely used in magnetic
recording technology, materials science and microelectronics. MFM is derived
from atomic force microscopy, which was invented by Binnig, Gerber and
Weibel in 1986 (Binnig et ai., 1986). It was able to measure the forces
between a sharp tip and a surface using a cantilever. Extending this idea, in
1987, Martin and Wickramasinghe developed MFM that placed a magnetic tip
on the cantilever and used a heterodyne interferometer detector to study the
stray fields of magnetic structures in magnetic materials (Martin and
Wickramasinghe, 1987).
MFM uses a cantilever probe containing a magnetic tip, which is
oscillated near its resonant frequency and scanned laterally at a constant height
above the sample. The variation in the magnetic interaction at different
regions of the sample changes the resonance properties of the cantilever and is
manifested as an amplitude change, frequency shift, or phase change. The
magnetostatic force is long-ranged and the magnetostatic force gradient is
detected in MFM. In general, the force gradient is proportional to the second
spatial derivatives of the individual components of the magnetic field
emanating from the sample. Major recent developments focus on the
quantification of MFM images, improvement of the resolution and the
application of external fields during measurements. One of the important
components for achieving these goals is the improvement of the sensing probe,
i. e., the magnetic tip.
There are many review articles about the principle and the application of
magnetic force microscopy (Grutter et al., 1992; Porthun et al., 1998 ;
Gomez, 2001). In this chapter, we first describe only the principle of
magnetic force microscopy from the view point of what determines the contrast
of the images. We show a few improvements in the fabrication of magnetic
force microscopy tips in our laboratory: CD high magnetic coercivity (He)
CoPt MFM tips for imaging domains under an applied magnetic field;
2 Sy-Hwang Liou

(2) superparamagnetic and low stray magnetic field MFM tips for imaging
domains of soft magnetic materials; @ electron-beam-induced-deposited
(EBID) MFM tips; point-dipole MFM tips; @ focused ion beam milled MFM
tips for high resolution magnetic domain imaging.

1. 2 Magnetic Force Microscopy

1. 2. 1 Basic Operating Principles of MFM


A brief overview of the principles of MFM is given here. Figure 1.1 shows the
schematic representation of the basic principles of MFM.
Deflection Laser beam
sensor

M""ri'----- t~/
Stray tip ~tilever
field ""............. . . .
l :. . . \/ V~_...:.

Sample -1-1-1-1-
Figure 1. 1 Schematic representation of the basic principles of MFM.

An MFM system consists of a flexible cantilever suspended from one end.


On the free end, a small magnetic tip is mounted. When a magnetic sample is
close to the tip, the tip interacts with the stray magnetic field emanating from
the sample. The interaction between tip and sample can be measured by a
sensitive optical deflection detector, which is used to detect cantilever motion
and hence to measure the force gradient.

1. 2. 2 Force Gradient Detection


The long-range electrostatic and magnetostatic forces are measured by the
derivative of the force or the force gradient using alternating current (ac) slope
detection. The cantilever is vibrated normal to the surface. In the absence of
magnetic forces, the cantilever has a resonant frequency f o. This frequency is
shifted by an amount /1f proportional to vertical gradients in the magnetic
forces on the tip. The shifts in resonant frequency tend to be very small,
typically in the range 1 - 50 Hz for cantilevers having a resonant frequency
f o ,....,100 kHz. These frequency shifts can be detected three ways: phase
Advanced Magnetic Force Microscopy Tips for Imaging Domains 3

detection, which measures the cantilever's phase of oscillation relative to that


of the piezo drive; amplitude detection, which tracks variations in oscillation
amplitude; and frequency modulation, which directly tracks shifts in resonant
frequency (Digital Instruments instruction manual, 1998).
For the amplitude detection method, Martin et a!. (1987) have shown
that the sensitivity measurements would yield

I:!.A =- 20 I:!.F'A o
3~Ko
where A o is the cantilever amplitude, K o is the spring constant, 0 is the
quality factor of the cantilever oscillation response and I:!.F' is the force
gradient. The order of the ampl itude is 20 - 100 nm. A change in force
gradient causes a shift in the resonance curve. When the cantilever is driven at
a fixed frequency Wd' this results in a change in the oscillation amplitude I:!.A .
Similarly, the phase change 1:!.5 of the oscillation is related to the force
gradient as

I:!.o = OI:!.F'.
Ko
The phase curve decreases with increasing frequency and crosses the center-
line (corresponding to a 90 phase lag) at the peak frequency. The phase
curve then measures the phase lag between the drive voltage and the
cantilever response. Again, vertical gradients in the magnetic force cause a
shift I:!.WR in the resonant frequency. The frequency shifts give rise to phase
shifts I:!. 0 , which then can be used to give an image of the magnetic force
gradients.
The shift in resonant frequency can be detected directly with a frequency
modulation (FM) technique. In the FM method, the cantilever is oscillated
directly at its resonant frequency using a feedback amplifier with amplitude
control (Albrecht et al., 1991; Durig et al., 1992). The change in resonant
frequency can be directly detected by FM demodulation techniques.

I:!.WR
WR
~- 2K I:!.
F'
o
Phase detection and frequency modulation produce results that are generally
superior to amplitude detection, with greater ease of use, better signal-to-
noise ratios, and reduced artifact content. The force gradient, as small as
10- 6 N/m, can be easily detected (Martin et a!., 1987).

1. 2. 3 Sensitivity and Resolution


The lift scan height (LSH) determines the sensitivity and resolution. In
general, MFM resolution is roughly equal to the lift height (Digital Instruments
4 Sy-Hwang Liou

instruction manual, 1998). Smaller LSH gives better resolution. Conversely,


magnetic features smaller than the LSH may not be resolved. The tip also
experiences stronger fields close to the surface, giving improved signal-to-
noise ratios. However, the tip has to be far enough from the surface (larger
than 10 nm) to avoid Van der Waals forces that will map topography on the
magnetic signal.
Resolution also depends on the size of the magnetic volume of the tip.
Smaller tips will provide higher spatial resolution. However, the sensitivity
depends on the magnetic field strength of the tip. Thus, a large interacting
volume is required to get high sensitivity. In general, the sensitivity and
resolution should be maximized in the tip design. Currently, the tip size can be
reduced to less than 50 nm and forces as small as 10 - 18 N can be detected
(Stowe et aI., 1997). A lateral resolution of about 10 nm has been claimed
(Grutter et aI., 1990a), but a more typical lateral resolution is about 30 nm
(Skidmore and Dahlberg, 1997; Digital Instruments instruction manual, 1998;
Chou et aI., 1986).

1. 2. 4 Tip-Sample Interaction
Figure 1. 2 shows that different forces acting on a tip dominate at different
distance regions (Porthun et al., 1998). The magnetic force dominates for tip-
surface distances greater than 10 nm. The magnetic forces measured in the
MFM are purely magnetostatic, i. e., they arise from the magnetic dipoles in
the tip interacting with dipoles in the sample. Typically, the distance between
tip and sample in MFM measurement is 10- 100 nm.

.
Quantum mechanical forces
..
Capillary attraction

van der Waals forces

.
Electric and magnetic forces
~

I nm IOnm IOOnm lJlm


Tip sample distance

Figure 1. 2 Different forces acting on a magnetic tip and the distance region where they
dominate the MFM signal (Porthun et al., 1998).

The force acting on the tip is the negative of the gradient of the
magnetostatic energy:

F =- V E = ~of v V (Mt;p Hsamp,e)dV'liP

where M liP is the magnetic moment of the tip and H samp,e is the stray field from
the sample. In this equation, the mutual interaction is assumed to have no
Advanced Magnetic Force Microscopy Tips for Imaging Domains 5

effect on the magnetic properties of either the probe or the sample. The
components of the force gradient are obtained by differentiation with respect
to Xi'

The above two equations give expressions for the force and force gradients
along the given axis. In general, the cantilever will experience forces in all
three directions, and the gradients will contain mixed derivatives with respect
to the coordinate variables.
By considering some special cases, the equations of force and force
gradient can be simplified. For example if we assume that the cantilever
vibrates nominally along the z-axis, then only the z-component of the force
gradient is sampled. Further, the cantilever is constrained to deflect only
along the z-axis, which restricts the measured force to be only the z-
component. In this case, the equations reduce to (GrLitter et aI., 1992):

Fz =lJo ~
""
;
f v
.
M / 3Hid
'1
Oz
V' =lJo. f( V
Hx
M x 33z + My Hy
33z + Mz 33Hz) d V'
z

F'
z. z
= ""
/
f
M 3 Hid V' =
lJo ~ V. / 3 Z2 lJo
2
f V.
(M 3 Hx
x 3 Z2
2 2
+ M Y 33 Z2
Hy + M z
2
3 Hz) d V'
3 Z2

These two equations show that the force and force gradients are mixtures of
the contributions of the z derivatives of the x, y, and z-components of the
surface field. The force is proportional to the first derivative, and the force
gradient is proportional to the second derivative of the individual components
of the field. The relative component of the stray field that is sensed depends
on the orientation of the magnetic moment of the tip. In practice, the tip is
magnetized along a definite axis. If the tip's magnetic dipole moment is along
the z-axis, magnetizations of x and y components of the tip are zero. The
force gradient becomes:

F 'z. z = IJ 0 f v
Mz 3
2

oZ
H zdV' .
---:::;-z

This equation has been widely used for the analysis of MFM contrast.
However, this relation is not enough for quantitative analysis. In most
situations, the magnetization of the tip is not uniform and exclusively along the
z-axis. Likewise, the condition of oscillation being purely along the z-axis is
seldom satisfied. In fact, owing to instrument design restrictions, the tip is
mounted at some angle a, typically 10 off the vertical axis. Thus, not only
the z-component, but also the x-component contributes to the MFM images.
To use the force gradient expressions to analyze MFM images, a precise
knowledge of tip magnetization is important. If the length of the tip domain is
6 Sy-Hwang Liou

much smaller than the characteristic dimensions of sample patterns, the tip can
be viewed as a point dipole (Hartmann, 1989), where the force is
proportional to the first derivative of the local field and the force gradient is
proportional to the second derivative. If the length of the tip domain is much
larger than the characteristic dimensions of sample patterns, the tip can be
thought of as a monopole tip or point charge tip (Schonenberger and Alvarado,
1990), where the force is proportional to the field and the force gradient is
proportional to the first derivative of the field. In general, higher gradients
give considerably narrower features (GrUtter et al., 1992). A point-dipole tip,
in principle, provides better resolution than a point-charge tip. But a true
point-dipole tip is not easily fabricated. The dipole approximation is only valid
as long as the separation between the charges in the tip is small in comparison
with the distance of observation (Rugar et al., 1990).
MFM images of longitudinal recording media give us an intuitive way to
understand the force gradient equation. In longitudinal media, the
magnetization is parallel to the sample surface. To record information, a
magnetic recording head is flown over the recording medium with a spacing of
several hundred nanometers or less. The recording process creates oppositely
magnetized regions with head-to-head or tail-to-tail magnetic transitions,
which generate a substantial stray field. It is the second derivative of this stray
field that is sensed by the MFM tip for imaging as shown in Fig. 1. 3. The
magnetization of tip is along the x-axis (top one) or z-axis (bottom one), thus
the x-component or z-component of the field derivatives was sensed. The
parallel field component has an odd symmetry about the transition, while the
perpendicular component has even symmetry (Rugar et al., 1990).

M~
MFMimage

0J~
MFM
trace

M of recording
media

MFM
trace

MFM image (b)

Iflm

Figure 1.3 Analysis of MFM images of recording media (a reference sample from NIST) .
(a) The magnetization of the tip was parallel to the film surface; (b) The magnetization of
the tip was perpendicular to the film surface. The white dot is a topographic feature of the
recording medium that indicates the same position of the magnetic images at (a) and (b).
Advanced Magnetic Force Microscopy Tips for Imaging Domains 7

1. 2. 5 Force Sensor
The most critical part of any force microscope is the force sensor, i. e., the
combination of a sharp tip on a flexible cantilever. In MFM, the measured
interaction forces are long ranged and quite small (10- 13 to 10- 18 N) (Martin
et al., 1987; Stowe et al., 1997). This defines some of the necessary
properties for tips suitable for MFM. The optimum value of the cantilever
spring constant is a trade off between force sensitivity and minimum tip-to-
sample spacing. Softer cantilevers enhance sensitivity, but are more
susceptible to crashing when the force gradient exceeds the spring constant.
The most widely used MFM cantilevers have spring constants in the range of
O. 1- 10 N/m. Another desirable property of force sensors is high resonant
frequency, which renders the cantilever insensitive to noise and vibrations. In
order to achieve a low spring constant and a high resonant frequency
simultaneously, the cantilever should have low mass. This can be achieved by
making the cantilevers physically small. Typical cantilevers are a few hundred
micrometers in length and about 25 IJm in width with a thickness in the range of
O. 1 - 10 IJm. To achieve high sensitivity to magnetic interactions, the tip
should have a magnetic moment large enough for the interaction to be
detectable with a good signal-to-noise ratio. At the same time, a tip with a
large magnetic moment will have an influence on the sample. To keep the tip's
influence on the sample as small as possible the spatial extent of the effective
magnetic tip volume has to be kept as small as possible, thereby minimizing
the spatial extent of the long-range dipolar magnetic tip-sample interaction and
maximizing spatial resolution in MFM. High resolution can be achieved if the
tip has a small radius of curvature and a small cone angle. Stray fields from
samples can be strong enough to change the magnetization of the tip. This will
alter the imaging mechanism during the measurement. The tip should be made
magnetically hard so that it stays stable in the applied fields. The above
parameters should be optimized simultaneously. Which type of tip is used
depends on the sample to be measured.
There are many methods to fabricate MFM tips. Some MFM force sensors
have been made by electrochemically etching thin ferromagnetic foils or wires
(Mamin et aI., 1988). The materials that have been used include iron, nickel
and cobalt wires, amorphous FeBSiC wires and nickel foil. Nickel is a
particularly good material to use because of its good etching properties and
corrosion resistance. Now, batch fabricated tips made by coating integrated
tips of silicon or silicon nitride with thin magnetic films, such as Co, CoPtCr or
NiFe are widely used and are commercially available (Gri.itter et aI., 1992;
Rugar et aI., 1990; Wolter et aI., 1991). Magnetic thin film tips have the
substantial advantage of a significantly reduced tip stray field as compared to
bulk wire tips, which is important when imaging magnetically soft samples
8 Sy-Hwang Liou

such as Permalloy. Another advantage of thin film tips is that choosing


appropriate coating materials can control the magnetic properties of tips.
Further improving the resolution of a tip is possible by reducing the physical
size of the magnetic volume of the tip. Several groups have reported the EBID
tips (Skidmore and Dahlberg, 1997; Fischer et aI., 1993; Ruhrig et aI.,
1994). EBID technique combines thin film and electron beam fabricated tips.
EBID carbon needles have been grown onto the end of the tip in a scanning
electron microscopy (SEM) chamber using a well-known contamination
technique (Ruhrig et aI., 1994; Broers et al., 1976). G. D. Skidmore's group
reported the EBID spikes could be grown with diameters averaging 30 nm and
tapering down to 5 nm (Skidmore and Dahlberg, 1997).

1. 3 Development of Advanced MFM Tips

There are many ways to improve the resolution of magnetic imagines. One of
the routes is through the development of advanced tips for MFM
measurements. We describe here a few of the tip innovations that we have
developed in our laboratory.

1. 3.1 High Coercivity CoPt MFM Tips

If the magnetization direction of a MFM tip is changed during measurement,


then the magnetic images are difficult to interpret. This is also a problem in the
determination of the magnetization direction of the MFM tip, when the
demagnetization field cannot be neglected due to the shape of the tip. These
problems can be resolved using high He MFM tips because their He is higher
than the demagnetization field. The main characteristic of the high coerciveity
magnetic force microscopy tip is that the direction of the magnetization is fixed
if the coercive field of the tip is higher than the magnetic stray field and any
magnetic field appl ied to the sample. The high coercivity magnetic force
microscopy tips are useful for the domain imaging of samples with high
magnetic stray field (such as permanent magnets, recording heads, etc. ) and
for domain imaging under an applied magnetic field.
1. 3. 1. 1 Enhancement of Coercivity in Nanometer-Size CoPt Crystallites
To fabricate high coercivity MFM tips, we need to find a suitable alloy with the
desired properties. The CoPt binary alloy is an excellent system because of its
chemical stability and high magnetic anisotropy. The anisotropy of CoPt
compounds is as high as 4 x 107 erg/cm3 (1 erg = 10- 7 J) and the saturation
Advanced Magnetic Force Microscopy Tips for Imaging Domains 9

magnetization is about 800 emu I cm 3 As shown in Fig. 1. 4, we have


prepared CoPt thin films with a thickness of 5 nm by dc magnetron sputtering.
After annealing in an Ar/H 2 atmosphere at a temperature of 650C for 12 h, we
showed that a magnetic coercivity (He) of 20 kOe was obtained in a CoPt thin
film that contains separated nanometer-size CoPt crystallites. From atomic
force microscopy and magnetic force mic~oscopy studies, the magnetic single
domain size of CoPt is in the range of 100 to 200 nm. The high He is due to the
well-separated nanometer-size crystallites and the well-ordered FCT phase of
CoPt alloy. A magnetic coercivity as high as 37 kOe in CoPt films can be
achieved by proper annealing (Liou et aI., 1999).

:i
(a) (b)

800
600 T=300k
snm
~ 400
200 II

~g-200o.
~ -400
-----~-~:I- Annealed at
6S0C
-600 for 12 h
-800 L........L_--'-----'_--'--_"------'-_-'--
-60 -40 -20 0 20 40 60
H(kOe)
(c)

Figure 1. 4 (a) The atomic force microscopy image of a 5 nm-thick film contains well
separated nanometer-size crystallites in the range of 100 nm to 400 nm. The height of
crystallites is in the range of 20 to 80 nm. (b) The MFM image was obtained using a CoPt
MFM tip magnetized parallel to the sample surface. The light and dark contrast
corresponds to the strength of the stray-field gradient on the sample surface. The lighter
color represents frequency shift in the MFM tip when the magnetization of the sample and
that of the MFM tip are repulsive. The crystallites with one light and dark area are single-
domain (as indicated by "S"); the grains that may contain a few crystallites with two or
more light and dark areas are multi-domain (as indicated by "M"). The size of a single-
domain crystallite Is between 100-200 nm. (c) An He value of 20 kOe and a saturation
magnetization of 668 emu/cm 3 were observed in the sample annealed at 650C for 12 h.
10 Sy-Hwang Liou

L 3. L 2 Tip Preparation
The tips were fabricated by deposition of a 30 nm thick CoPt film on
commercially available batch fabricated, micromachined cantilevers. Tip
fabrication involves three steps: magnetic thin film deposition, thermal
treatment and tip magnetization. The films were DC magnetron-sputtered with
the base pressure of 2 x 10- 8 Torr. The argon pressure during sputtering was
15 mTorr. The temperature in the chamber was 150 C. A CoPt target was
used to sputter the magnetic layer, and the deposition rate was 20 nm/min.
The as-deposited tips were annealed in a furnace under a flow of a gas mixture
of 20% hydrogen and 80% argon. After annealing at 650C for 12 h, the films
were found to have FCT structure with a coercivity in excess of 20 kOe (Liou
et ai., 1999). The tips were magnetized in a field of 7 T using a
superconducting magnet. The magnetic field was applied along a direction 10
(off the vertical axis (that is, perpendicular to the sample surface).
Figure. 1.5 shows a typical high He CoPt MFM tip with a 100 nm size
magnetic particle at the end of the tip.

Figure 1. 5 A high-coercivity CoPt MFM tip. A 100 nm size magnetic particle was formed.
at the end of the tip.

We demonstrated that the He of these MFM tips is higher than the gap
field of a typical thin-film recording head (Liou and Yao, 1998). As shown in
Fig. 1. 6, the magnetic transitions and magnetic grains in a recording disk
were clearly observed. This is due to the better defined magnetization
direction and very small size of the high He CoPt MFM tip. An example of using
the high He tips in the study of magnetic domain images in the presence of an
external magnetic field is shown in Fig. 1.7. In this study, the magnetic field
was applied along the hard axis [110J of a patterned epitaxial [110J Fe film.
The sample has a 10 IJm width, 200 IJm length, and 100 nm thickness. The
magnetization curve was measured by a SQUID magnetometer. The detailed
domain evolution can clearly be seen, showing that we are able to directly
correlate the magnetic domain patterns with the magnetization curve.
Advanced Magnetic Force Microscopy Tips for Imaging Domains 11

Figure 1. 6 An MFM image of a thin film hard disk (a reference sample from NIST) using
a high He CoPt MFM tip which was magnetized perpendicular to the thin film surface.

2.5
_[110]
2.0

1.5

1.0
~
u
u
......
~ ~

~
.a,
0.5
(,
il,1
/';.'
.
M
O 0
2S Vl
-0.5
::E
-1.0

-1.5

-2.0
-2.5 L----L-_L--~~~~=__5~~~~'......L___.J
-500 -400 -300 -200 -100 0 100 200 300 400 500
Applied field (Oe)

Figure 1. 7 Magnetic domain structures in a patterned epitaxial (110) Fe film under a


magnetic field. The film was magnetized at - 5, 000 Oe before the measurement. (a) At
- 400 Oe, the edge domain wall is formed, (b) between- 300 and - 270 Oe the domain
wall propagates from the edge into the sample. It shows a rapid change of magnetization.
(c) At - 260 Oe the domain structure is formed in the film. (c) - CD Between - 260 Oe
and 260 Oe, the domain walls only move a short distance. This shows flux closure
domains. (g) At 300 Oe the domain wall starts to disappear and domain walls exist only at
the edge and at the surface (the surface domain wall has a much lower signal, i. e., light
color). (h) At 400 Oe, the sample is nearly saturated, however, the surface domain wall
is still visible. In contrast, there is no surface domain in the domain pattern (a) (that was
magnetized at - 5, 000 Oe before the measurement). The surface domain disappears at
an applied magnetic field of about 600 Oe. The CoPt MFM tip was magnetized
perpendicular to the thin film surface.
12 Sy-Hwang Liou

Figure 1.8 shows a theoretical loop prediction based on the assumption of


moving 180 walls. There are domains along both [110 J and [100 J type
directions. For iron, the [100 J directions are easy directions, where as the
energy of the [110 J type domains is enhanced by K 1 V/4, where K 1 =
0.05 MJ/m3 is the first anisotropy constant and V is the volume of the [110J
domain. However, from the angular dependence of the cubic anisotropy it
follows that both [11 OJ and [100J are minima, that is, they are separated by
an energy maximum and an intermediate in-plane angle. This explains why the
magnetization does not simply rotate in an applied field but prefers to reverse
by domain-wall motion. In zero field (case I), there is no net magnetization in
the [110 J direction, while the size of the domains is determined by the
competition between bulk anisotropy and domain-wall energy. In an applied
field, the domains in field direction grow at the expense of the anti parallel
domains (II) until there are no antiparallel domains left (III). At this point,
where H = Hs ' there is a singularity, and a complicated cooperative process
turns the magnetization in a common [110J direction (IV). The mathematical
analysis of the process shows that the magnetization jump from III to IV
depends on the first anisotropy constant K 1 and on the (average) domain size.
On the other hand, the slope of the M (H) can be used to calculate the energy
of the domain walls. These considerations yield the right orders of magnitude
for K 1. The remaining differences may be ascribed to residual rotation
processes and to the fact that the domains are not in perfect equilibrium.

H
Figure 1. 8 Theoretical loop and domain interpretation.

Starting from the completely saturated state, H = + 00, reverse domains


are nucleated in the vicinity of H s . The direct observation of nucleation
processes is one of the most demanding problems in magnetism. Figure 1.9
shows that the nucleation starts at the edges of the bar and expands into the
center of the bar. This process is accompanied by a variety of magnetic
viscosity phenomena.
Advanced Magnetic Force Microscopy Tips for Imaging Domains 13

(b)

10 IlJII
--t'"

(c) (d)

Figure 1. 9 Nucleation of reverse domains. These are the details of magnetic domain
structures of a patterned epitaxial [11 OJ Fe film under a magnetic field ranging from 290 to
270 Oe. It clearly shows that the domain wall starts to propagate from the edge into the
sample. There is a rapid change of magnetization from (a) 290 Oe. (b) 280 Oe.
(c) 275 Oe. to (d) 270 Oe.

1. 3. 2 Superparamagnetic and Low Stray Magnetic


Field MFM Tips
To image magnetic domains of soft magnetic materials (low magnetic coercive
field materials such as garnets) we need some special precautions due to the
fact that the domain structures of soft magnetic materials are easily perturbed
by the stray field from the magnetic force microscopy tips. We have
developed two types of tip coatings, the superparamagnetic and low stray
magnetic field tips, for the magnetic domain imaging of soft magnetic
materials.
The first type of tip coating developed ("soft" MFM tips) is based on
superparamagnetic Fe particles embedded in a Si02 matrix. Details of the film
preparation and its magnetic properties can be found in previous publications
(Liou and Chien, 1991; Kanai and Charap, 1991; Malhortra et al., 1994). A
principle feature of superparamagnetism is the lack of hysteresis; that is, it
has zero remanance and magnetic coercivity. The Fe particles can rotate
freely in the presence of the stray magnetic field from the sample, and thus
cannot cause the sample magnetization to reverse (which has been a severe
problem with conventional "soft" MFM tips). In contrast to the assumed fixed
magnetic moment, M, of ferromagnetic coated tips, M for superparamagnetic
tips depends on the applied magnetic field. The force between the tip and the
sample is always attractive, so that interpretation of the images is relatively
straightforward (similar to Bitter techniques). Because of the field
dependence of the magnetization of the tip, the interaction is very sensitive to
the strength of the magnetic field as well as the field gradient.
14 Sy.Hwang Liou

As shown in Fig. 1. 10, the magnetic domain images of a garnet film are
clearly observed. There is no observable interference of the domain pattern of
the garnet by the MFM tip. The field-gradient profile is symmetric, which is
expected due to the superparamagnetic character of the tip. The tips used in
this experiment were commercial silicon micromachined cantilevers coated
with superparamagnetic Fe70 (Si0 2 )30 films. The cantilevers have spring
constants of 1 - 5 N/ m, resonant frequencies of 70 - 89 kHz, and qual ity
factors of about 200 in air. The magnetic garnet is an epitaxial YGdTmGa/
YSmTmGa film, grown on a nonmagnetic Gd3Ga5 0 12 (GGG) substrate, with a
zero-field stripe width of 1.4 IJm, that is similar to that in the references (Katti
et aI., 1992; Wadas et aI., 1994; Tian et aI., 1997). The superparamagnetic
MFM tip was used to obtain high resolution imaging of magnetic recording
heads (Liou et aI., 1997>.

3Hz

OHz

Figure 1. 10 A magnetic domain image of a garnet using a superparamagnetic MFM tip.

The second type of tip is a high coercivity, low stray field MFM tip that
made the influence of magnetic tips on the sample negligible. The coating
developed is based on permanent magnetic materials, such as CoPt alloys
with a reduced magnetization volume that lowers the emitted magnetic stray
field from the tip. The coating procedure is similar to that of high coercivity
CoPt MFM tips with a smaller coating thickness. As shown in Fig. 1. 11, the
magnetic domain image of a garnet was obtained using MFM with low magnetic
stray field and high coercivity MFM tips. The CoPt MFM tip was fabricated by
deposition of 7 nm of CoPt. The MFM tip was magnetized perpendicular to the
film surface.
From the domain width of these two pictures, these two types of MFM tips
do not appear to disturb the domain structure of the garnet. The newly
developed superparamagnetic and high coercivity with low stray field MFM tips
show great promise for magnetic domain images of soft magnetic materials.
Advanced Magnetic Force Microscopy Tips for Imaging Domains 15

La

0.5

Figure 1. 11 A magnetic domain image of a garnet using a high coercivity and low stray
magnetic field MFM tip.

1. 3. 3 Electron-Beam-Induced-Deposited (EBID) MFM Tips

A typical EBID MFM tip is shown in Fig. 1. 12. As can be seen, the spike
tapers down its 300 nm length and has a diameter of about 70 nm at the end of
the tip. The EBID tip has a finger shape that allows a particle to form at the
end of the tip. The size of the disk-like particle is about 80 nm in diameter and
20 nm in thickness. As can be seen from the SEM picture, the other magnetic
particles are a few hundred nm away from the tip. This difference in the tip
shape and the distribution of the magnetic particles allows a better spatial
resolution and smaller stray field of the EBID MFM tip than that of a typical thin
film MFM tip.

Figure 1. 12 An electron-beam-induced-deposited MFM tip (it has a high coercivity and a


very small magnetically active area at the end of the tip) .
16 Sy-Hwang Liou

As shown in Fig. 1. 13, we compared domain images of a Co/Pt


multilayer film with ion irradiation using thin film and EBID MFM tips. Ion beam
irradiation has been shown to locally alter the magnetic properties of Co/Pt
multilayer films. The magnetic film was patterned by irradiated N+ ions with a
dose of 10 16 ions/cm2 at 700 keY through a silicon stencil mask having about
1. 3 Ilm diameter holes. Localized modification of the magnetic coercivity and
easy magnetization axis with ion irradiation was recently demonstrated by
Chappert et al. (1998). At a high enough dose, 10 16 ions/cm2 , the easy axis
of Co/Pt multilayer films is rotated into the plane of the film. As shown in Fig.
1. 13a, the magnetic force microscopy image has a roughly circular pattern
that is similar to what is expected for a 90 domain wall. Figure 1. 13a is the
MFM image of an irradiated area of a Co/Pt multilayer using a thin film MFM
tip (with 20 nm CoPt coating). The MFM contrast is very different inside and
outside of the irradiated area. The irradiated area (i. e., soft magnetic region)
was influenced by the stray field from the tip. It shows a dark spot in the
roughly circular irradiated region. The measured result, as shown in Fig.
1. 13a, is similar to that of the reference <Terris et aI., 1999). In order to
reduce the interference of the tip with the sample, we have used a low stray
field and high coercivity tip (an EBID MFM tip). Figure 1. 13b is the domain
image of irradiated Co/Pt multi layers on the same sample using an electron
beam modified MFM tip with a 20 nm CoPt coating. This magnetic image is
very different from that of the images obtained by the thin film coated tip. The
MFM image shows essentially no tip perturbation and the MFM contrast is very
similar inside and outside of the irradiated area. The magnetic domain
structure before irradiation is granular with a magnetic grain size (about 100
nm) that is a few times larger than the topography grain size (about 30 nm)
obtained from AFM. The magnetic domain structure after irradiation is also

(a) (b)

Figure 1. 13 Magnetic images of an inhomogeneous sample, a Co/Pt multilayer (10


periods of 3 A Co and 10 A Pt deposited on Si), patterned by irradiated W at
700 keV through a silicon stencil mask having about 1. 3 ~m diameter holes. (a) used
a thin film coated MFM tip; (b) used an electron beam induced deposition MFM tip.
The MFM tips were magnetized perpendicular to the film surface.
Advanced Magnetic Force Microscopy Tips for Imaging Domains 17

granular with a magnetic grain size similar to that of the non-irradiated region.
However, the domain structure at the transition region between the irradiated
and non-irradiated regions of the sample has a strong contrast (the 90 domain
wall). The 90 domain wall is pinned by the non-irradiated high perpendicular
anisotropy region. The full width at half maximum response over the 90
domain boundary is about 70 nm. We showed that the MFM image using an
EBID MFM tip is much less intrusive than that of a thin film MFM tip. The EDIB
MFM tip with low stray magnetic field is suitable for the magnetic domain
studies of both soft and hard magnetic materials.

1. 3. 4 Point-Dipole MFM Tips


MFM images are usually modeled as a monopole or dipole in which the
magnetic charges near the tip point are assumed to be constant (Grutter
et aI., 1992; Hug et aI., 1996). However, the magnetic interaction is a long-
range interaction that includes not only the interaction between the sample and
the magnetic material near the tip, but also from the extended area of the tip.
This makes it more difficult to interpret MFM images. One of the ways to
improve the resolution of MFM images and faith in MFM images is to improve
the magnetic tips used by developing tip coating techniques or by modifying
tip-shape (Ruhrig et aI., 1994; Liou and Yao, 1998; Grutter et aI., 1990b;
Hopkins et aI., 1996; Shearwood et aI., 1996). In this section, we compare
the magnetic domain images obtained by using a point MFM tip and a thin-film
MFM tip. The image obtained from the point tip shows much fewer effects due
to the stray field from the extended area of the tip. The spatial resolution of
the MFM images using the point tip is clearly better than that of the thin-film
tip. The tips used in this experiment were commercial silicon micromachined
cantilevers with spring constants of 1 - 5 N/ m, resonant frequencies of 70 -
89 kHz, and quality factors of about 200 in air. The thin film MFM tips were
made by magnetron sputtering of an amorphous metal of Fe6] CO l8 B 14 Si
(2605CO). The thickness of the coating is 30 nm. The amorphous 2605CO
alloy is a soft magnetic material that has been used as the core of high power
transformers. One of the unique features of amorphous alloy is its structure and
magnetic isotropy. The domain structure of an amorphous magnetic material is
expected to be less sensitive to the shape of the tip than a crystalline magnetic
material. The magnetization of the amorphous alloy, 1. 75 T, is about two
. times higher than that of Permalloy. The high magnetization is essential for the
improvement of the signal-to-noise ratio. In general, the smaller tip magnetic
moment will improve the spatial resolution and make it easy to interpret MFM
images, but may demand improved signal-to-noise. The coercive field of these
tips was about 20 Oe, smaller than the typical stray field of hard disk recording
media.
The point MFM tips were prepared as follows: Q) The magnetic thin film
tip was mounted on a scanning probe microscope. The photo-resist film was
18 Sy-Hwang Liou

painted on the glass slide. ~ The tip was gradually dipped into the photo-
resist film by controlling the force on the tip. The coating size on the tip can be
easily controlled by the thickness of the photo-resist film, which was about
O. 5 - 1 IJm. @ The unprotected magnetic film was then removed by ion
sputtering performed in a vacuum chamber. The photo-resist on the tip was
removed by acetone. As shown in Fig. 1. 14, an MFM cantilever with a
magnet size of 500 nm on the tip is clearly observed.

Figure 1.14 An MFM cantilever with a magnet size of 500 nm on the tip CLiou. 1999b).

The sample investigated was a reference sample made from a recording


disk typical of those currently found in hard-disk drives. The reference sample
has been characterized by Rice et al. (1997). One of the advantages of using
the reference sample is that the same area can be easily located and
compared.
We compared the images obtained from the thin film tip and the point tip.
Both tips are coated with the same magnetic material and show a soft magnetic
behavior (i. e., the tip magnetic moment follows the stray field of the
sample). As shown in the top of Fig. 1. 15, these atomic force microscopy
images, taken from a reference sample, are from the same area. For a soft
MFM tip. the interaction of the stray field from the recording disk with the tip'
generates a magnetic force derivative that is always attractive (as shown by
the dark lines). The bottom of Fig. 1. 15a, b shows the MFM images that
were obtained from the thin-film tip. The bottom of Fig. 1. 15c is the MFM
image that was obtained from the point tip. The difference between
Fig 1. 15a, b is that the tip was magnetized in an opposite direction. The
magnetization direction of the tip was perpendicular to the sample surface.
Figures 1. 15a, b show the effects of the magnetic interactions from an
extended area of the tips. At further distances from the sample, which has a'
smaller stray field, the tip magnetization does not flip completely. As a result,
the tri-bit pattern (indicated by an arrow) was not resolved, as shown in
Fig 1. 15a, b. We observed either one dark line or two dark lines for the tri-bit
pattern using a thin film tip that has a different magnetization direction. In
contrast, Fig 1. 15c shows three dark lines for the tri-bit pattern using a point
tip. This can be better illustrated by Fig. 1. 16. The insets of Fig. 1. 16 are
Advanced Magnetic Force Microscopy Tips for Imaging Domains 19

the schematic of the magnetization of the MFM tip near the tri-bit pattern. It
illustrates that the magnetic interactions between the tip and the sample can be
compl icated. because the magnetic hysteresis in the MFM tips causes history-
dependent interactions between the tip and the sample.

r 5 Mm! / :
~JlliI

(a) (b) (el

Figure 1. 15 A comparison of the MFM images with the thin-film tip (a) and (b) and the
point tip (c). The details are explained in the text (Liou, 1999).

EttEJ
C0\[\JCOY
N
:B -1.0
~~.O
g -3.0
N
:B -1.0
~~~
g -3.0
'N'
C 0
:B -1.0
~ -2.0
g -3.0
g. -4.0 g. -4.0 g.-4.0
J: -5.0 ~O-::-'-::-----.,.'-:---,-'-::---::-'-:-
0.5 1.0 1.5 2.0
J: -5.0 ~O-::-'-::-----.,.'-:---,-'-::---::-'-:-
0.5 1.0 1.5 2.0
J: -5.0 ~O-::-'-::-----"''-:---,-'-::---::-'-:
0.5 1.0 1.5 2.0
Length (11m) Length (11m) Length (11m)
(a) (b) (e)

Figure 1. 16 Frequency shifts on 2 Mm line traces across a tri-bit pattern on the reference
sample with the thin-film tip (Fig. 1. 15a, b) and the point tip (Fig 1. 15c). The scale of
the frequency shift at Fig. 1. 15c is 5 times less than that of Fig. 1. 15a, b. The inset is
the schematic drawing of the tip magnetization. The actual magnetization direction of the
tip is not known. Due to the shape anisotropy of the film at the tip, it is expected that have
a component other than z-direction (Liou, 1999).
20 Sy-Hwang Liou

Figure 1. 16a, b are the 2 IJm trace across the tri-bit area from the thin-
film tip. Figure 1. 16c is the 2 IJm trace across the tri-bit area from the point
tip. The traces in Fig. 1. 16a, b show that the tri-bit patterns were not
resolved. There are either two peaks or one peak, depending on the
magnetization direction of the tip. It is clearly shown in the insets of
Fig. 1. 16a, b that the magnetic interaction between the stray field of the
sample and the magnetization directions at the side-wall of the tip is the key
factor for the difference. In contrast, Fig. 1. 16c shows the magnetization of
the tip as it follows the stray field of the sample, possessing three peaks, with
essentially no effect due to the magnetization interaction from the side-wall of
the tip. However, the sensitivity of the point tip is about 5 times less than that
of the thin film tips. This comparison clearly shows that the point tip produced
a better-resolved MFM image because it was not affected by the magnetic
interactions from the extended area of the tip.

1. 3. 5 Focused Ion Beam Milling MFM Tips


We have successfully fabricated a very sharp MFM tip. Figure 1.17 shows an
SEM image of a very small MFM tip made by focused ion beam milling. The tip
was first coated with a magnetic film then machined by a focused ion beam
source, so that there is a nanometer-size magnetic particle on the very end of
the tip. The smaller size of the magnetic particle results in much improved
lateral resolution. Figure 1. 18 shows the domain configuration of a 150 nm-
thick epitaxial [110J Fe film obtained by an MFM tip that was machined by a
focused ion beam source.

Figure 1.17 A CoPt MFM cantilever machined by a focused ion beam source.
Advanced Magnetic Force Microscopy Tips for Imaging Domains 21

Figure 1. 18 The domain configuration of a 150 nm-thick epitaxial [11 OJ Fe film obtained
by an MFM tip that was machined by a focused ion beam source.

The image was obtained with a vertically magnetized tip in a zero applied
magnetic field to the sample. As shown in the above magnetic images the
MFM tip innovations undoubtedly improved imaging capability. The arrows
indicate the magnetization direction. The full width at half maximum of the
Bloch domain wall width was measured to be 60 - 70 nm, which agrees well
with the calculated value for that of bulk Fe. Figure 1. 19 is the MFM signal
across a Bloch wall. The sol id line is a theoretical fit that incorporates lowest-
order (bulk) magnetocrystall ine, magnetoelastic, surface, and magnetostatic
contributions. The above MFM tip innovations undoubtedly improved imaging
capability. We will continue a systematic investigation of this advanced tip
technology.

1000

500

o ......... . . ....7 .-~-...-....-...-....-...-i


..

::8 -500
~
-1000
V
-1500

0 200 400 600 800


x(nm)

Figure 1. 19 MFM signal across a Bloch wall. The solid line is a theoretical fit that
incorporates lowest-order (bulk) magnetocrystalline, magnetoelastic, surface, and
magnetostatic contributions.
22 Sy-Hwang Liou

References
Albrecht, T. R., P. GrLitter, D. Horne, D. Rugar. J. Appl. Phys. 69: 668
( 1991)
Binnig, G., C. F. Quate, C. Gerber. Phys. Rev. Lett. 56: 930 (1986)
Broers, A. N., W. W. Molzen, J. J. Cuomo and N. D. Wittels. Appl. Phys.
Letts. 29:596 (1976)
Chappert, C., H. Bernas, J. Ferre, V. Kottler, J. -Po Jamet, Y. Chen, E.
Cambril, T. Devolder, F. Rousseaux, V. Mathet and H. Launois. Science
280: 1919(1998)
Chou, S. U., M. S. Wei and P. B. Fischer. IEEE Trans. on Mag. 30:4485
( 1986)
Digital Instruments, Dimension 3100 series scanning probe microscope
instruction manual, pp. 13 - 1 and 13 - 10, 1998
Durig, U., O. Zuger, A. Stalder. J. Appl. Phys. 72(5): 1778 (1992)
Fischer, P. B., M. S. Wei, and S. Y. Chou. J. Vac. Sci. Technol. B 11;
2570 (1993)
Gomez, R. D. In: Marc De Graef and Yimei Zhu eds. Experimental Methods
in the Physical Sciences, Vol. 36. Academic Press, London, pp. 69 - 143
(2001)
GrLitter, T., Th. Jung, H. Heinzelmann, A. Wadas, E. Meyer, H. -R. Hidber
and H. -J. GLintherodt. J. Appl. Phys. 67: 1427 (1990a)
GrLitter, P., D. Rugar, H. J. Mamin, G. Castillo, S. E. Lambert, C. J. Lin,
R. M. Valletta, O. Wolter, T. Bayer and J. Greschner. Appl. Phys. Lett.
57: 1820 (1990b)
GrLitter, P., H. J. Mamin and D. Rugar. Magnetic force microscope. In: R.
Wiesendanger and H. -J. GLinterodt eds. Scanning Tunneling Microscopy
II. Vol. 28. Spring-Verlag, Berlin, pp. 151-207(1992)
Hartmann, U. Phys. Lett. A 137: 475 (1989)
Hopkins, P. F., John Moreland, S. S. Malhotra and S. H. Liou. J. Appl.
Phys. 79:6448 (1996)
Hug, Hans J., B. Stiefel, A. Moser, I. Parashikov, A. Klicznik, D. Lipp,
H. -J. GLintherodt, Gabriel Bochi, D. I. Paul and R. C. O' Handley. J.
Appl. Phys. 79: 5609 (1996)
Kanai, Y. and S. H. Charap. J. Appl. Phys. 69: 4478 (1991)
Katti, R. R., P. Rice, J. C. Wu, H. L. Stadler. IEEE Trans. Magn. 28:
2913 (1992)
Liou, S. H. and C. L. Chien. Appl. Phys. Lett. 52: 512 (1991)
Liou, S. H., S. S. Malhotra, John Moreland and P. F. Hopkins, Appl. Phys.
Lett. 70: 135 (1997>
Liou, S. H., Y. D. Yao. J. Magn. Magn. Mater. 190: 130 (1998)
Liou, S. H., S. Huang, E. Klimek, R. D. Kirby, Y. D. Yao. J. Appl. Phys.
85: 4334 (1999)
Advanced Magnetic Force Microscopy Tips for Imaging Domains 23

Liou, S. H. IEEE Trans. on Magn. 35: 3989 (1999)


Malhortra, S. S., Y. Liu, J. X. Shen, S. H. Liou and D. J. Sellmyer. J.
Appl. Phys. 76: 6304 (1994)
Mamin, H. J., D. Rugar, J. E. Stern, B. D. Terris, S. E. Lambert. Appl.
Phys. Lett. 53: 1563 (1988)
Martin, Y., H. K. Wickramasinghe. Appl. Phys. Lett. 50: 1455 (1987)
Martin, Y., C. C. Williams and H. K. Wickramasinghe. J. Appl. Phys. 61:
4723 (1987)
Porthun, S., L. Abelmann, C. Lodder. J. Magn. Magn. Mater. 182: 238
( 1998)
Rice, P., Stephen E. Russek, Jay Hoinville and Michael H. Kelly. IEEE
Tran. Magn. 33: 4065 (1997)
Rugar, D., H. J. Mamin, P. Guethner, S. E. Lambert, J. E. Stern, I.
McFadyen, T. Yogi.J. Appl. Phys. 68: 1169 (1990)
Ruhrig, M., S. Porthun and J. C. Lodder. Rev. Sci. Instrum. 65: 3224
(1994 )
Schonenberger, C., S. F. Alvarado. Z. Phys. B 80: 373 (1990)
Shearwood, C., A. D. Mattingley, M. R. J. Gibbs. J. Magn. Magn. Mater.
162: 147 (1996). new entry Skidmore, G. D. and E. D. Dahlberg. Appl.
Phys. Lett. 71: 3293 (1997)
Stowe, T. D., K. Yasumura, T. W. Kenny, D. Botkin, K. Wago and D.
Rugar. Appl. Phys. Lett. 71: 288 (1997)
Terris, B. D., L. Folks, D. Weller, J. E. E. Baglin, A. J. Kellock, H.
Rothuizen and P. Vettiger. Appl. Phys. Lett. 75: 403 (1999)
Tian, Fang, Chen Wang, Guangyi Shang, Naixin Wang, Chunli Bai. J. Magn.
Magn. Mater. 171: 135 (1997)
Wadas, A., J. Moreland, P. Rice, R. R. Katti. Appl. Phys. Lett. 64: 156
(1994)
Wolter, 0., Th. Bayer and J. Greschner. J. Vac. Sci. Technol. B 9: 1353
( 1991)

The author thanks Jon Orloff, L. Gao. R. F. Sabiryanov. and Ralph Skomski for their
contributions. The author also thanks B. D. Terris. Liesl Folks. D. Weller. A. D. Kent.
John Moreland. Paul Rice and Steve Russek for the samples that were used in this study.
The Research was supported by Army Research Office. grant DAAD 19-00- 1-0 119 and
Nebraska Research Initiative grant.
2 Lorentz Microscopy and Holography
Characterization of Magnetic Materials

Daisuke Shindo. Young-Gil Park

2. 1 Introduction

In order to understand the properties of advanced magnetic materials, it is


necessary to make clear both their microstructure and magnetic domain
structure. Transmission electron microscopes can be utilized for
characterization in both microstructure and magnetic domain structure. In the
present chapter transmission electron microscopy for observing the magnetic
domain structure, i. e., Lorentz microscopy and electron holography are
presented. Some analytical methods which are useful for magnetic domain
structure analysis are also presented prior to Lorentz microscopy and electron
holography.
Before going to the explanation of each analysis or microscopy, a brief
review of the history in electron microscopy on magnetic domain structure is
given. Using an electron mirror microscope, magnetic domains were first
investigated by Mayer (1957). Subsequently, by using a defocusing method
(Fresnel mode) of a transmission electron microscope, Hale et al. (1959)
observed successfully magnetic domain walls. Since the principles of imaging
magnetic domain structure could be understood in terms of the Lorentz force on
electrons, electron microscopy for observing magnetic domains is known as
Lorentz microscopy. Lorentz microscopy with a conventional transmission
electron microscope consists of the Fresnel method (see Section 2.4.2. 1) and
the Foucault method (see Section 2.4.2.2). The latter method with the use of
an objective aperture was first utilized by Boersch et al. (1959). By using the
coherent beam with a field-emission gun, the Foucault method has been
extended to make the interference of some selected beams to image the
magnetization distribution (Coherent Foucault method, Chapman et al.,
1994). The method for imaging magnetization distribution using a split detector
and electron beam scanning, the so-called differential phase contrast Lorentz
scanning transmission electron microscopy (DPC Lorentz STEM) (see Section
2.4.3) was developed by Chapman arid Darlington (1974) to utilize the phase
contrast for imaging the detailed magnetization distribution.
On the other hand, holography was invented by Gabor (1949) for
Lorentz Microscopy and Holography Characterization of Magnetic Materials 25

enhancing the resolution of electron microscopes. Cohen (1967) noted for the
first time that the holographic technique could give direct information about
magnetic domain structures. Tonomura (1972) and Pozzi and Missiroli (1973)
subsequently demonstrated that the information about the domain structure
actually is reflected in the electron phase distribution. In 1980. the lines of
magnetic flux inside ferromagnetic particles were first observed as contour
fringes in interference micrographs by Tonomura et al. (1980). In the phase
reconstruction process. the sophisticated optical system has been utilized
(Matsumoto and Takahashi. 1970; Endo et al.. 1979). Recently. since
various film scanners and new recording systems such as the slow-scan CCD
camera and the imaging plates have been developed. the phase reconstruction
process is now widely carried out with the Fourier transform on the digital data
of the hologram (Tanji et al.. 1993; McCartney and Zhu. 1998; Park et al..
2000; also see Section 2. 5) .

2. 2 Instrumentation

In observation of the magnetic domain structure in magnetic materials by


transmission electron microscopy. special attentions should be paid to the
following two points:
(1) Magnetic field of the objective lens;
(2) Thin film effect.
The magnetic field at the specimen position in a transmission electron
microscope is about 1.5 T. and this strong magnetic field destroys or modifies
the inherent magnetic domain structure. Thus. the magnetic field at the
specimen position should be reduced especially for observing soft magnetic
materials. One of the easiest ways to reduce the magnetic field is to switch off
and degauss the objective lens. Usually by these processes. the residual
magnetic field can be reduced to less than 0.2 mT. However. fine
transmission electron microscope images are not expected under this
condition. since the objective lens. which determines the resolution of a
transmission electron microscope. is off. For keeping the action of the
objective lens for observing detailed magnetic structure. the position of the
specimen should be shifted. leaving the strong magnetic field or the special
shield for the magnetic field should be introduced in the objective lens. In the
latter case. for instance. there is a magnetic field free objective lens with a
single-gap or a double-gap proposed for observing the magnetic domain
structure. Figure 2. 1 shows an example of the objective lens of the double-gap
type (Tsuno and Inoue. 1984) which was used for DPC Lorentz STEM (see
Section 2.4.3). The pole-piece has two gaps of 8 1 and 8 2 among three pole-
pieces and a hole. The specimen is surrounded with the middle pole-piece and
26 Daisuke Shindo, Young-Gil Park

shielded from the magnetic field. Figure 2.2 illustrates typical examples of the
magnetic field distributions of the objective lens around the specimen position
and electron trajectories for various microscope modes (Tsuno, 1988). Being
different from conventional electron microscopy (Fig. 2. 2a, b), the strong
magnetic fields form below the specimen in Fig. 2. 2c and both sides of the film
specimen in Fig. 2. 2d.

Middle pole-tip

Lower pole-tip

Figure 2. 1 Cross section of the double gap objective lens pole-piece for Lorentz
microscopy.

For observing transmission electron microscope images, specimens


should be thin, and thus the magnetic domain structure is considered to be
basically different from that in bulk materials. Actually, it is considered that in
thin fi Ims less than a few tens nm in thickness, the Neel walls are observed in
perwalloys, while in rather thick films more than several tens nm in thickness,
the Bloch walls tend to appear (Neel, 1955). Between such thicknesses, the
cross-tie walls appear (Huber et a!., 1958; Tonomura, 1999). Thus, the
thickness of the specimen should be as large as possible. From this point of
view, the usage of a high-voltage transmission electron microscope is useful,
since relatively thick specimens can be observed with high energy electrons.
Also, it should be noted that thickness evaluation is a very important job for
understanding the magnetic domain structure observed by transmission electron
microscopy. In Section 2. 3. 2, the method of evaluating specimen thickness
with analytical electron microscopy will be presented. Finally, it should be
pointed out that in thin films the effect of the demagnetizing field perpendicular
to the film surface is notable. In order to appropriately analyze the magnetic
domain structure by transmission electron microscopy, the effect of the
demagnetizing field should be carefully taken into account.
Lorentz Microscopy and Holography Characterization of Magnetic Materials 27

Detector
(a) (b)

(c) (d)

Figure 2.2 Schematic illustration of magnetic field distributions of the objective lens and
electron trajectory. (a) Conventional transmission electron microscopy; (b) Conventional
STEM; (c) Lorentz microscopy; (d) Lorentz STEM.

In the following study on magnetic domain structure with Lorentz


microscopy where the observation of the magnetic domain structure and the
evaluation of the domain wall width were performed, two types of electron
microscopes were used. One is a high-voltage electron microscope (JEM-ARM
1250) with an accelerating voltage of 1250 kV. This is basically a high-
resolution electron microscope, but since the position of the specimen in the
high-voltage electron microscope can be shifted to a higher position above the
objective lens, the magnetic field at this position is reduced to about 0.5 mT.
A 300 kV transmission electron microscope (JEM-3000F) was also used for
observation of magnetic domain structures. In the case of the 300 kV
transmission electron microscope, it is usually hard to observe the magnetic
domain structure of soft magnetic materials, since the magnetic field of the
objective lens weakly excited at the position of the specimen for magnetic
28 Daisuke Shindo, Young-Gil Park

structure observation is still about 0.5 T. Thus, before observing the magnetic
domain structure in soft magnetic materials, the magnetic field around the
specimen was reduced to around 0.2 mT by switching off and then degaussing
the objective lens. Lorentz microscope images were recorded with not only
conventional EM films but also the slow-scan CCO camera (Mooney et al.,
1990; Shindo and Hiraga, 1998) and the imaging plates (Sonoda et aI., 1983;
Taniyama et aI., 1997), having good linearity between the electron intensity
of transmission electron microscope images and the output signals.
Electron holography for the magnetic structure characterization shown was
carried out with a 200 kV transmission electron microscope (JEM-2010F) and a
300 kV transmission electron microscope (JEM-3000F), the latter of which
was also used for observation of Lorentz microscope images as noted above.
Both microscopes were installed with thermal field-emission guns (FEGs). As
shown in Table 2.1, like a cold FEG (Tonomura, 1992), a thermal FEG
provides the highly coherent electron beam which is necessary for electron
holography study. In the 300 kV transmission electron microscope, an electron
energy-loss spectrometer is also attached. Each microscope with a thermal
FEG is equipped with a biprism. A biprism consisting of a fine Pt filament of
about O. 6 IJm in diameter and two earth potential electrodes are positioned
between the objective lens and the intermediate lens (Shindo et al., 2000). In
other electron microscope systems, the diameters of the biprism are reported
to be 0.3 - 0.5 IJm (Joy et aI., 1993; Tanji et aI., 1993). In order to adjust the
width of interference fringes appropriately, the biprism can also be set
between the 1st intermediate lens and the 2nd intermediate lens in other
electron microscope systems (Frost et aI., 1995). This configuration is useful
for adjusting the magnification, especially when the objective lens is switched
off. In the following study with electron holography, the magnification of
images was set to be about 4000 times and the exposure time for taking
electron holograms was about 8 s. The spacing of the interference fringes was
in the range of 18 - 25 nm and about 210 interference fringes were observed in
the electron hologram. Electron holograms observed with conventional EM
films were digitized using a film scanner with a resolution of 8 IJm/pixel.

Table 2. 1 Comparison in characteristics of various electron gunsC200 kV) .

Thermionic emission Field emission

Characteristic Thermal FEG Cold FEG


W LaB 6
ZrO/W"OOl W(lOO) W 010l

Brightness at 200 kV
CA/ Ccm 2 sr
-S X 10 5 -sx 106 -S X 10 8 -S X 10 8 -S X 10 8

Source size SO IJm 10IJm O.l-llJm 10-100 nm 10-100 nm


Energy spreadCeV) 2.3 1. S 0.6-0 8 06-0.8 0.3-0. S
Lorentz Microscopy and Holography Characterization of Magnetic Materials 29

2. 3 Analytical Electron Microscopy for Structure


Characterization

2. 3. 1 Outline of Current Analytical Electron Microscopy


Currently, not only the resolution of electron microscopes themselves, but also
the performances of their peripheral instruments are greatly developed.
Among the peripheral instruments, electron energy-loss spectrometers and
energy dispersive X-ray spectrometers have been utilized extensively. Also, it
is noted that electron guns of field-emission types are introduced in analytical
electron microscopes. Being different from the electron guns of a thermionic
type, i. e.,a tungsten hairpin type and a LaBs point filament type which have
been used widely, the electron guns of field-emission types provide high beam
intensity and small probe size. Also, the latter guns provide the high coherency of
the electron beam which is necessary for electron holography study.
In the following sections, two standard analytical methods, i. e., electron
energy-loss spectroscopy (EELS) and energy dispersive X-ray spectroscopy
(EDS) will be presented. By using EELS, thickness evaluation can be easily
carried out (Section 2.3.2). On the other hand, compositional analysis which
is important for microstructure analysis can be effectively performed by the
EDS elemental mapping method as shown in Section 2. 3. 3.
Principles of EELS and EDS can be understood in terms of one of the
inelastic electron scattering processes, i. e., the excitation of core electrons.
Figure 2.3 shows the change of electronic structure due to the excitation of the
1 s inner shell electron and the resultant energy-loss spectrum and X-ray
spectrum observed. Consider the case where the incident electron gives such
energy to the specimen with which the electron in the 1 s orbital (K-shell) is
excited to higher energy level. Since the energy levels below the Fermi level
are all occupied by electrons in the ground state, the electron is excited to the
unoccupied state above the Fermi level. Thus, when the incident electrons
lose the energy I:1E larger than the energy difference between the 1s level and
the Fermi level, the probability of the inelastic scattering increases
drastically, and eventually the sharp peak appears at the energy I:1E in the
energy-loss spectra. Since the threshold energy at the edge I:1E is specific to
each material the specimen can be identified with this energy value.
Figure 2.4 shows electron energy-loss spectra in a Ndu Fen B 18 . 5
nanocomposite magnet, where B-K edge can be detected. Although the signal
is rather weak, it can be used for distinguishing the Fe3 B grains from ex - Fe
grains (Park et al., 2001). Furthermore, from the threshold energy of I:1E and
30 Daisuke Shindo, YoungGiI Park

the shape of the edge, the information of atomic bonding or electronic structure
can be obtained (Egerton, 1996). In addition, there are other excitation
processes of atomic electrons, such as collective excitation of valence
electrons (so-called plasmon excitation) and the transition between the energy
bands. It should be noted that most of the inelastic scattering process
appearing in the energy-loss spectra consists of the plasmon excitation.

L3 - - - - - - 4 > - - 1 f -
L2 - - - - + - - - - - - - j f -

K -----+-----<0--

E E(=hv)
(a) (b)

Figure 2.3 Schematic illustration showing the change of electronic structure due to the
excitation of the 1s inner shell electron and the resultant energy-loss spectrum (a) and X-
ray spectrum (b) observed.

On the other hand, some surplus energy will be emitted as the


characteristic X-ray or the Auger electron, when the atom in the excited state
comes down to the ground state. In the former case, the electron in the higher
energy level fills the hole in the lower energy level in a manner satisfying the
selection rule. Like EELS, the energy of the characteristic X-ray can be used
for identifying constituent elements, and also from the integrated intensity of a
peak in the spectrum, the composition of the material can be evaluated. In the
schematic illustration of Fig. 2. 3b, the characteristic X-ray emission process
in the transition from L3 to K, and the characteristic X-ray emitted being called
KCXl are shown. There are also several other characteristic X-rays, which are
frequently used for compositional analysis, such as K131 and Lcx! corresponding
to the transitions from M3 to K and from Ms to L3 , respectively.
Lorentz Microscopy and Holography Characterization of Magnetic Materials 31

50 nm

_-B

2000.---....---------------,

1500
,-,
::i
~
o
.iii
1000

~
500

o 160 200 240 280


Energy loss,E(eV)

Figure 2.4 Electron energy-loss spectra obtained from areas A and B in the bright-field
image (upper part) of a Ndu Fen B 18 . 5 nanocomposite magnet annealed at 993 K.

2.3.2 Thickness Measurement by EELS


There are several methods for measuring thickness with a transmission
electron microscope. such as the contamination spot-separation method, the
extinction contour method and the convergent beam electron diffraction
(CBED). Each has its own limitation in accuracy and lor application as briefly
noted below.
Contamination spot-separation method: This was first used by Lorimer
et al. (1976), and it is generally used, if the contamination can be generated.
However, this method has relatively large errors from -- 50 % to -- 200 %
(Stenton et ai., 1981).
Extinction contour method (Hirsch et ai., 1977): This method makes it
simple to evaluate the specimen thickness at anyone point by counting the
number of thickness fringes from the edge of the crystal (Rogulic, 1964). In
this method, errors in the specimen thickness measurement occur mainly due
to the deviation from the exact two-beam diffraction condition.
32 Daisuke Shindo, Young-Gil Park

Convergent beam electron diffraction method: By measuring the spacing


of the Kossel-Mbllenstedt fringes in convergent beam diffraction patterns
(Kelly et aI., 1975; Allen, 1981), the specimen thickness for crystalline
materials is determined. This method is the most popular and accurate
method, but is useful only for materials with good crystall inity.
Comparing with these three methods, EELS is an easy and useful method
of evaluating thickness. As shown in Fig. 2.5, an electron energy-loss
spectrum mainly consists of a zero-loss peak and a plasmon peak which
corresponds to the elastic and inelastic scattering, respectively. In the so-
called log-ratio method in EELS (Egerton, 1996), the specimen thickness t is
determined by the following relationship:

J.- = In(!2) (2.1)


i\ p 10
where i\ p is the mean free path for inelastic electron scattering, lois the
integrated intensity for elastically scattered electrons (so-called zero-loss
peak) in a spectrum and IT is the whole intensity. It is noted here that i\ p for
each material should be known for determining the absolute specimen
thickness with Eq. (2. 1). For crystalline materials, the CBED method can be
uti! ized for evaluation of specimen thickness (Allen, 1981). When the
thickness is evaluated by the CBED method, the mean free path i\ p can be
obtained by measuring the energy-loss spectra at the same position. Once i\ p
is obtained, the thickness of the specimen can be easily evaluated from EELS.
While the CBED method needs a good crystalline region, the accuracy in
thickness measurement with EELS does not basically depend on the
crystallinity. Actually, taking into account i\ p for Fe, the thickness of the
Ndz Fe14 B magnet has been estimated to evaluate the magnetic flux density
(X 105)
12

10 /zero-Ioss peak
~

:::i 8
~
0 6
Vi
c
~ 4

2 Plasmon peak

0
o 50 100 150 200
Energy loss, E (eV)

Figure 2.5 Electron energy-loss spectrum obtained from NdzFe" B.


Lorentz Microscopy and Holography Characterization of Magnetic Materials 33

(Zhu and McCartney, 1998; Park et ai., 2002). It should also be noted that in
electron holography the reconstructed phase image includes the information of
both thickness changes and the magnetic flux. Thus, in order to extract the
information on the magnetic flux from the reconstructed phase image, the
thickness should be obtained with EELS (see Section 2.5.3. 1) .

2. 3. 3 Elemental Mapping with EDS

Energy dispersive X-ray spectroscopy (EDS) is useful for quantitative


compositional analysis. Since the electron probe obtained with an FEG can be
minimized to a diameter less than 1 nm, local compositional analysis can be
effectively carried out. By scanning the probe on the specimen surface and
detecting the characteristic X-ray intensities, the elemental mapping can be
carried out. This is contrasted with the elemental mapping with EELS as noted
in the next chapter. The elemental mapping with EELS is especially useful for
light elements, while that with EDS is a quantitative method, efficient for
heavy elements. The elemental mapping or compositional analysis in a local
area is important not only for understanding the microstructure in magnetic
materials but also the evaluation of thickness measurement with EELS, since
the evaluation is sensitive to the accuracy of the mean free path for inelastic
electron scattering which depends on the composition.
In the following, the elemental mapping for Sm-Co permanent magnets for
clarifying the distribution of additives is demonstrated. It is well known that the
properties of the Sm-Co magnets are very sensitive to the amount of the
additives and the heat treatment. In order to analyze the distribution of the
additives, EDS elemental mapping was carried out with a probe of O. 7 nm.
Images at the top left in Fig. 2. 6a, b show scanning transmission electron
microscope images of 2 h-isothermal-aged with coercive forces of 40 kA/m and
step-aged with 836 kA/m Sm-Co magnets, respectively. In both Fig. 2.6a,b,
characteristic X-ray images of the elements Cu, Fe and Zr are presented. As
indicated in the figures, both magnets containing a small amount of additives
such as Cu, Fe and Zr, consist of matrix phases of 2 : 17 Rand 1 : 5 H, and a
precipitate phase, the So-called z phase. It is clearly seen that the additives
are highly segregated by the step aging. Especially, the segregation of Cu at
the cross points (arrowhead) in Fig. 2.6b in domains of the 1 : 5 H phase is
noted. Also, Zr is clearly located at the z phase with a thickness of about
1nm. These results indicate that the step aging treatment leads to the chemical
partitioning among the phases without any microstructural changes, resulting in
the high coercive force (Mishra et al., 1981; Shindo et al., 2001).
34 Daisuke Shindo, Young-Gil Park

(a)

(b)

Figure 2.6 Characteristic X-ray images showing elemental distributions in 2 h-isothermal-


aged Ca) and step-aged Cb) 8m-Co magnets.
Lorentz Microscopy and Holography Characterization of Magnetic Materials 35

2. 4 Lorentz Microscopy on Magnetic Domain Structure

2. 4. 1 Principles of Lorentz Microscopy


The Lorentz force F for an electron passing through a specimen with the
magnetic flux density B is given by
F =- e( v x B) (2.2)
where v is the electron velocity, e is the elementary electric charge. The
direction of the force expressed by the vector product in Eq. (2.2) is given by
Fleming's left hand rule. Thus, the Lorentz force on an electron passing
through a magnetized specimen, which has the magnitude of magnetization L s '
is given by
F=-e(vxIs) (2.3)
Here, the direction of the magnetization in the specimen of thickness t z is
assumed to be normal to the plane of diagram (y direction) as shown in
Fig. 2.7. By the magnetization Is, the electron is deflected from the straight
trajectory, but the energy of the electron remains constant. From Eq. (2.3),
the Lorentz force on the incident electron being parallel to the x direction is
expressed as
Z
dP x
- = m -d= x evl (2.4)
dt d tZ Y

where m is the electron mass at the velocity v,


Z
m = me/( 1 - 2v t )
(2.5)

with me being the rest mass and c is the velocity of the light. In Eq. (2.4),

d~x is the x component of the Lorentz force, giving P x = m ~~ = etzl y , since


the velocity of the electron can be treated as a constant. For the magnitude of
magnetization given by I y = I s and I x = I z = 0, the deflection angle is given
by

e = dx = dx s!! = et z I (2.6)
L dz d t dz mv y'

The deflection angle by the Lorentz force is rather small compared to Bragg
angles. For instance, the deflection angle (e L ) in Ndz Fe14 B with thickness of
100 nm and I = 1.6 T is 3 X 10- 5 rad. for 1250 keV electrons.
36 Daisuke Shindo, Young-Gil Park

Incident beam

Figure 2. 7 Schematic illustration showing the deflection of electron beam by a foil


magnetized in the y direction.

The magnetic deflection or the phase shift of an electron wave can be


utilized to image magnetic domains with several different ways. In the
following, we first explain Lorentz microscopy with conventional transmission
electron microscopes, i. e . the Fresnel method and the Foucault method
(Hirsch et aI., 1977; Reimer, 1993).

2. 4. 2 Lorentz Microscopy Using Conventional Transmission


Electron Microscopes

2.4.2. 1 Fresnel Method (Defocusing Method)


The simplest domain formation is one consisting of 180 domains in a uniaxial
0

material. Here we assume that the 180 domains are magnetized


0

perpendicular to the plane of the diagram and in opposite directions in


alternate domains. According to Fleming's left hand rule, the electron beam is
deflected in opposite directions in adjacent domains, leading to a deficiency of
electrons and an excess of electrons due to overlap in the region below the
lower specimen surface. If we observe the specimen on the overfocus
condition, the increase and the decrease of the electron intensity are observed
at the positions of the domain walls periodically as shown in Fig. 2. 8b. On the
other hand, if we observe the specimen on the underfocus condition, the
reversed image contrast appears as shown in Fig. 2. 8c. A typical example of
Lorentz microscope images of an as-sintered 8m-Co magnet observed by the
Fresnel mode is shown in Fig. 2.9. Under the just focus condition, there is no
definite image contrast, while under the defocused condition, white or black
Lorentz Microscopy and Holography Characterization of Magnetic Materials 37

lines appear at the positions of magnetic domain walls depending on the sign of
defocus. The width of the white or black lines directly depends on the focus
setting. Taking into account the experimental condition such as focus setting
carefully, the wall width can be evaluated quantitatively as shown below.
Incident beam

IIIIAAAA\\\\
t tt IIIIIIII
!11/flAtt"" JIll/I!!
IIIIIIAAIIIIII 11111111
Under IIIIIIIAII\\\\\ 1111111I
focus

Over
focus

(a)

(b)
,..
I \
I \
I \
-, /
\ /
\ I

(c)
'/
Figure 2.8 (a) Schematic ill,ustration showing Lorentz microscopy with the Fresnel mode,
Image intensities on overfocus condition (b)and underfocus condition (c),

The domain wall width in magnetic materials can be determined by a


linear extrapolation method proposed by Reimer and Kappert (1969).
According to the geometric-optical theory (Kappert, 1970), a full width W d
(1lf) at half maximum (FWHM) of a divergent wall image at a defocus value
M is given by (Kappert and Schmiesing, 1971)

Wd (llf) = 2Xl/2(M) +2M9 L (J)Y<;1/2) (2.7)

where (J) is the magnetic flux inside the specimen, (J) Y (x) is the y component
of the magnetic flux, and 9L is the Lorentz deflection angle. The coordinates
x, y and z are indicated in Fig. 2. lOa. In Fig. 2. lOb, the wall width 25 is
defined by the slope of a function of d(J) Y (x) / dx at the wall center, whereas
2Xl/2 is given as FWHM of the function d(J)y (x)/dx as shown in Fig. 2.10c.
Therefore, by extrapolating FWHMs (Wd) of divergent wall images taken at
38 Daisuke Shindo, Young-Gil Park

(b)

(c)

Figure 2.9 Lorentz microscope images of an as-sintered 8m-Co magnet observed by the
Fresnel mode. (a)Just focus; (b)Underfocus; (c)Overfocus.

various defocus values, the domain wall width at 1).f = 0 can be experimentally
determined. Here, it is reported that the wall width obtained in this way was
shown to be not significantly influenced by an illumination angle a of the
incident beam, provided that a is of the order of 8L or less (Suzuki and
Hubert, 1970; Suzuki et al., 1984). Figure 2. 11 shows a series of Lorentz
microscope images observed by the Fresnel mode in an as-sintered Sm-Co
magnet. The images were observed systematically by changing the defocus
value. The wall widths were measured from the intensity profiles of divergent
wall images recorded on imaging plates. These intensity profiles were
obtained in the narrowest areas of divergent wall images to avoid
overestimation of the wall width because of tilting of the domain wall against
the electron beam. The wall widths plotted against the defocus values are on a
straight line as shown in Fig. 2.12, and eventually the domain wall width of an
Lorentz Microscopy and Holography Characterization of Magnetic Materials 39

as-sintered 8m-Co magnet is evaluated to be 10 nm (Yang et al., 1998). The


value is consistent with the assumption for evaluation of the coercive force of a
8m-Co magnet (Perkins et al., 1975; Nagel, 1979).
28

(a)

iP;,/<P
1 /
/
/
I.

-8 x

7
/
/
/ -I
(b)

x
(c)

Figure 2.10 (a) Schematic illustration showing a 180 domain wall in a film with uniaxial
anisotropy; (b)Wall width 26 defined by the slope of a function of d<l>y(x)/dx at the wall
center; (c) 2Xl/2 given as FWHM of a function d<l>y (x) /dx.

Figure 2. 13 shows the comparison of the intensity distributions of the


domain wall images in an as-sintered 8m-Co magnet observed by the Fresnel
mode with the calculated intensity distributions. In Fig. 2. 13, the squares and
solid lines indicate the observed and theoretical intensity distributions at the
domain wall. Consider the spin distribution across the wall region for the
minimum total energy resulting from exchange and either anisotropy or
magnetostatic energy (Kittel, 1949), the equation of the theoretical intensity
40 Daisuke Shindo, Young-Gil Park

(a) (b) (c) (d)

a-g
Under focus

(e) (f) (g)

(a') (b') (c') (d')

a'-g'
Over focus

(e') (f') (g')

Figure 2. 11 Variation of domain wall images of an as-sintered 8m-Co magnet observed


as a function of defocus value 1'.'. Images observed on underfocus condition are indicated
by (a) - (g) , whi Ie overfocus condition by (a') - (g') .

distribution of the domain wall images (Fuller and Hale, 1960;Wade, 1966) is
derived as
I ( U) I I = I 1 R sech2 u 1-1, U = u R tanh u (2.8)

where u = x I {j, U = X I {j, and R = f;.ffhl {j. {j is half


of domain wall
width, x and X are the coordinates normal to the wall direction in the
specimen and image planes, respectively. While I ( U) is the electron beam
intensity in the image plane, I is the uniform illuminating intensity. The plus
Lorentz Microscopy and Holography Characterization of Magnetic Materials 41

160

E
"
~. 120
~

"
OJ)

.'"
100

~ 80
~
4-<
0
60
~

~
w.. 40

20

0
-2000 -1500 -1000 -500 0 500 1000 1500 2000
Defocus value, D./(~m)

Figure 2. 12 FWHM of divergent wall images of an as-sintered 8m-Co magnet observed


as a function of defocus value t;.f.

and minus signs in Eq. (2. 8) refer to the divergent and convergent cases,
respectively. Equation (2. 8) indicates that the width of white lines in the
convergent images is significantly narrower than that of black lines in the
divergent images. As shown in Fig. 2.13, the experimental intensity profiles
measured with the imaging plate at a defocus value O. 63 mm were best fitted
with the calculated intensity profiles of the divergent and convergent cases at
R = 0.3 (Yang et aI., 1997).

1.6

1.4
~ 0.96 ~
S
'=:;' 0.88 ~ 1.2
0.80
0.72 1.0 00 o
0.64
0.56 '-------'---_-----'-,--_-L-_---'---_-----'-_ 0.8 '--'-_ _---'---_ _.L..-_---'-_ _--'--
-4 -2 0 2 4 -4 -2 0 2 4
U=X/J U=X/8
(a) (b)

Figure 2. 13 Comparison of intensity distributions calculated (solid lines) and observed


with the Fresnel method (squares). (a) Divergent case; (b) Convergent case. Defocus
value is 0.63 mm.
42 Daisuke Shindo, Young-Gil Park

2. 4. 2. 2 Foucault Method On-focus Method)


In the diffraction pattern obtained from an area containing a domain wall
separating two magnetic domains such as shown in Fig. 2. 14, the diffraction
spots split into two with an angular separation 2 BL . This splitting can be
observed directly on diffraction patterns, both in the transmitted and diffracted
beams. In Fig. 2. 15a, an electron diffraction pattern obtained from a sintered
NdzFe14 B magnet shows the split in each spot. The separation of the spots is
-5 X 10- 5 rad. Being consistent with the estimation noted in Section 2.4.1.
By displacing the objective aperture in the back focal plane, it is possible to
exclude one of the two transmitted beams. As a result, alternate magnetic
domains appear to be dark and bright regions periodically as shown in
Fig. 2. 14. Figure 2. 15b,c show a typical example of the magnetic domains of
a Ndz Fe14 B magnet observed by the Foucault mode.
Incident beam Incident beam Incident beam

Figure 2. 14
D [(J m
Schematic illustration showing Lorentz microscopy with the Foucault mode.

Figure 2. 16a, b show Lorentz microscope images of 2 h-isothermal-aged


and step-aged Sm-Co magnets observed with the Fresnel mode and the
Foucault mode. The images observed with the Foucault modes are inset. In
the image of Fig. 2. 16a, the magnetic domain walls are rather straight, while
serrated domain walls indicating the pinning at the cell boundaries are
observed in Fig. 2. 16b. From the images obtained by the Fresnel method, the
positions of the domain walls are clearly visual ized, but it is hard to observe
the detailed magnetic domain structure corresponding to the microstructure due
to the defocus condition. The Lorentz microscope images observed by the Foucault
method in Fig. 2. 16b clearly show that the 1 : 5 H phases of the cell boundaries act
as the attractive pinning center for magnetic domain walls in a step-aged magnet.
Combining EDS elemental mapping Csee Section 2. 3. 3) and Lorentz microscopy
with the Foucault mode, it is found that the domain wall pinning is greatly increased
with the increase of domain wall energy gradient between the 1 : 5 Hand 2 : 17 R
phases due to the chemical partitioning of the Cu and Fe atoms by step aging
CMishra et aI., 1981).
Lorentz Microscopy and Holography Characterization of Magnetic Materials 43

(a)

(b) (c)

Figure 2. 15 (a) Electron diffraction pattern of a sintered Nd2 Fe"B magnet; (b), (c)
Lorentz microscope images observed with the objective aperture eliminating one of the
splitting transmitted beams.

(a) (b)

Figure 2. 16 Lorentz microscope images of 2 h-isothermal-aged (a) and step-aged


(b) magnets. The images were obtained by the Fresnel method while the images inset
were obtained by the Foucault method.

On the other hand. the modified observation mode in the Foucault method
was developed (Chapman et al.. 1994). In this method. by eliminating half of
the diffraction pattern with a phase-shifting aperture. the interference contrast
in the domain can be obtained. Apertures themselves are chosen to be of the
44 Daisuke Shindo, Young-Gil Park

opaque half-plane, phase-shifting half-plane, or phase-shifting small hole kind.


The method is called the coherent Foucault method. If the edge of the phase-
shifting aperture is put in the center of a transmitted spot, the interference
contrast is observed in all directions. The coherent Foucault image is obtained
in real time without any post process. Therefore, the method is useful for an in
situ experiment (McVitie et aI., 1995).

2.4.3 Lorentz Microscopy Using Scanning Transmission


Electron Microscopes
Like electron holography presented in the following section, differential phase
contrast (DPC) Lorentz microscopy introduced by Chapman et al. (1978)
provides detailed magnetization distribution in magnetic materials. The
principle of DPC Lorentz STEM is as follows (Tsuno, 1988).
The electron beam condensed by the objective pre field lens is focused on
the magnetic specimen under scanning mode. At each point, due to the
deflection of the beam by the Lorentz force, two disc patterns are produced in
the back-focal plane of the objective lens (Fig. 2. 17). Usually, two discs are
crossed over each other, because the illumination angle at the specimen plane
is large compared with the deflection angle. With intermediate and projector
lenses, the discs on the objective back-focal plane are magnified at the split
detector (Fig. 2.18). When the convergent beam passes the left side domain
in Fig. 2. 17, deflection disc is shifted to the left side of the detector and the
signal detected by the detector (A + D) is stronger than that detected by the
detector (B + C) and vice versa. So, if the signal detected by (B + C) is
subtracted from that detected by (A + D), a component of the signal due to the
phase shift is enhanced and a component of the signal which is not produced by
the phase shift is vanished, because the latter component is detected equally
by both detectors. In this way, the magnetization direction can be detected.
Figure 2. 19 shows a schematic diagram of a quadrant detector imaging system
(Morrison et aI., 1980; Tsuno et aI., 1989). By adding or subtracting each
signal, domains with every direction can be displayed without rotating the
specimen as indicated in the right side of Fig. 2.19.
Figure 2.20 compares the magnetic domain images of a Co poly-crystal
taken by DPC Lorentz STEM (a) and by Lorentz microscopy with the Foucault
mode (b). The advantage of DPC Lorentz STEM over usual Lorentz
microscopy is noted. The image contrast change due to the specimen
thickness variation observed in (b) is removed in (a). Image of slip bands in
the horizontal direction also disappears by a careful adjustment of the detector
position CTsuno and Inoue, 1984). It is clearly seen that the image contrast
corresponding to magnetic domain structure can be extracted in the DPC
Lorentz STEM.
Lorentz Microscopy and Holography Characterization of Magnetic Materials 45

z~ \
Y@----x \
\

/
I
/
I

Lens,~/==tt==4====t=t:==i/>

Split detector

Figure 2. 17 Schematic illustration of electron ray diagram showing DPC Lorentz STEM
for a magnetic film with a 180 domain wall.

A B

D C

Figure 2. 18 Schematic illustration showing the discs corresponding to the splitting beams
arrowed on a quadrant detector.

A+B+C+D
(A+B)-(C+D) @
(A+D)-(B+C) @
I
I
A-C

I
I
B-D 0
'----rl D D (C-D):
L ~L ~

Quadrant Pre Main Signal Image


detector amplifier amplifier mixing selector

Figure 2. 19 A quadrant detector and the image processing system of DPC Lorentz STEM.
46 Daisuke Shindo, Young-Gil Park

(a)

Figure 2.20 A comparison of images of a Co poly-crystal taken by OPC Lorentz STEM


(a) and Lorentz TEM with the Foucault mode (b). The photographs are kindly provided by
Dr. K. Tsuno.

2. 5 Principles and Application of Electron Holography

2. 5. 1 Principles of Electron Holography

Among various electron microscopy techniques, electron holography provides


a unique method for detecting the phase shift of the electron wave due to the
magnetic field and electric field. Here, the phase shift means the phase
Lorentz Microscopy and Holography Characterization of Magnetic Materials 47

change relative to the electron plane wave travel ing in the vacuum,
exp[i(kz - wt) ] where
21T
k =-, w = 21TV (2.9)
A
where v is frequency. As shown in Fig. 2.21, electron holography is carried
out through the two step imaging process. In the first step, a hologram is
formed by interfering an object wave and a reference wave using a biprism. In
the second step, the phase shift is extracted from the hologram by using the
Fourier transform.
Field emission gun

Specimen

Biprism

Digital data

Electron
t

D~
hologram

Interference fringes

Reconstructed _------,1
phase image
-:::::==~~~J

Figure 2. 21 Schematic illustration showing the two step imaging process in electron
holography

Figure 2.22 shows a geometrical configuration for forming a hologram in


an electron microscope. An electron beam emitted from a field-emission tip is
accelerated and then collimated to illuminate an object through a condenser
lens system. An object is located in one half of the object plane being
illuminated with a collimated electron beam. Assuming that the object is
illuminated by a plane wave of a unit amplitude having a wave vector parallel
to the optical axis, the change of the scattering amplitude of the plane wave
due to the object is in general described as
48 Daisuke Shindo, Young-Gil Park

q(r) = A(r)exp[i(r)] (2. 10)

FEG

Condenser lens

Specimen

Objective aperture

Hologram

Figure 2.22 Geometrical configuration for forming a hologram.

where A (r) and (r) are real functions and describe the amplitude change
and the phase shift due to the object, respectively. Since specimens are
usually thin films, in most cases the vector r is confined in the film plane. Due
to the voltage of a biprism, the object wave and the reference wave are tilted

by ~h and - ~h, respectively. Thus, the scattering amplitude resulting from


the interference of the object wave and reference wave is given by

9h(r) = A(r)exP(-lTi ~hX + i(r))+ eXP(lTi ~hX). (2. 11)

The intensity of the hologram is given by

'h(r) =I 9h(r) 1
2
= 1 + A 2 (r) + 2A(r)Cos[2lT ~hX + (r) l (2. 12)

According to Eq. (2. 12), it is seen that the period (:h) of the interference
fringes in the hologram is modulated by the phase change (r) due to the
object. By performing the Fourier transform (21) of the hologram, one can
obtain
Lorentz Microscopy and Holography Characterization of Magnetic Materials 49

Y[/h(r)] =5(u) +Y[A 2 (r)]

+ Y[A(r)exp(- i(r] 5( ~h) * U -

+ Y[A(r)exp(i(r] * 5( + ~h) U (2.13)

where * is Convolution operation. 1st and 2nd terms are called


"autocorrelation" (corresponding to the central area in Fig. 2. 23c) . The other
two terms are called "sidebands." It is seen that the holographic information
about the phase shift and the amplitude change is reserved at the sidebands in
the 3rd and 4th terms of the right hand side in Eq. (2.13). Now, by selecting

the 4th term, shifting it by - ~h and performing the inverse Fourier transform
(y- 1 ) on this term, we obtain

y- I [yeA (r )exp( i (r] * 5 (u)] = A (r)exp[i (r)]. (2.14)

Thus, we can get the phase shift and the amplitude change in the reconstructed
image as digital data. In the following part, the intensity of the reconstructed
phase image Iph(r) is represented by cosine function, i. e.,
Iph(r) = cos[(r)]. (2.15)

If it is necessary, the amplification of the phase can be done by multiplying the


phase (r) by an integer n.
Now we consider the phase shift due to the electric potential cp and the
vector potential A as follows. In general, the phase shift due to these
potentials is given by

(r) =: (CPdt-AdS) (2. 16)

where e is the elementary electric charge, Ii is Planck's constant divided by


2rr, and the integration goes over any closed circuit in space-time (Aharonov
and Bohm, 1959). When t (time) is constant and the film is uniform in
thickness as shown in Fig. 2. 24, phase shift (r) is mainly produced by the
vector potential as given by

(2.17)

where A is vector potential. We evaluate the phase difference of the electron


wave at C and 0 due to the magnetic flux inside the specimen, where the
electron wave has the same phase at A and B. The phase difference between
C and 0 is given by

(rc) - (ro) =- ~fcA ds + ~fo A ds = ~,h A. ds. (2.18)


Ii A Ii B Ii JABDC

From the definition of the vector potential,


50 Daisuke Shindo, Young-Gil Park

(a)

(c) (d)

Figure 2. 23 ( a) Lorentz microscope image of a 8m-Co based permanent magnet.


Domain walls are indicate by arrows; (b) Electron hologram at the same area as (a);
(c) Digital diffractogram obtained from an electron hologram (b); (d) Reconstructed phase
image obtained by the inverse Fourier transform with the scattering amplitude in the circle
of (c).

A B
r---------------------- -------- -------------------1
I I
I
I
I
I
I
I
I

l-J
I
I
<P [B] I
--......._--+---:._--+----> I
I
I
I
I
I
I
I
L ~ ~
I

C D
Figure 2.24 Geometrical configuration of a magnetized thin film and an electron beam for
understanding the phase shift due to the magnetization.

B = rotA (2.19)
and using Stokes' theorem, the following relation is obtained

~A ds = ffrotAdS = ff BndS. (2.20)

Thus, we find

(2.21)

Here, (/) is the magnetic flux going through and being normal to the area
ABDC. If the phase difference between rc and rD is 2n,
Lorentz Microscopy and Holography Characterization of Magnetic Materials 51

!!.- (/) = 2n (2.22)


h
then,

(/) = ..!!- = 4. 1 X 10- 15 (Wb) . (2.23)


e

This is just twice the flux quantum (2he)' In this way, we have the relation
between the magnetic flux inside the specimen and the width l which
corresponds to the phase difference 2n, i. e. ,

(/) = It,B = eh (2.24)

where we assumed the constant magnetic flux density B inside the specimen.
Thus, the distance corresponding to the phase difference 2n is

(2.25)

Finally, we present the relation,


cp(rc) - cp(rD) = 2n (2.26)
where rc - rD =1. In the reconstructed phase images such as being presented
by cosine function, l is just the distance between white lines or black lines.

2.5.2 Practice of Electron Holography


Here, we present an example of hologram and reconstructed phase image
comparing with the Lorentz microscope image obtained with the Fresnel
method. Figure 2. 23a shows a Lorentz microscope image of a 8m-Co based
permanent magnet. The Lorentz microscope image of Fig. 2. 23a shows
magnetic domain wall as white and black bands indicated by arrows. An
electron hologram at the same area as Fig. 2. 23a is shown in Fig. 2. 23b. Due
to the strong magnetic field of this material, the interference fringes curve from
place to place in Fig. 2. 23b. Figure 2. 23c shows a digital diffractogram
obtained from the electron hologram of Fig. 2.23b. The bright regions in the
upper part circled and lower part of Fig. 2. 23c correspond to the 3rd and 4th
terms of Eq. (2. 13), respectively. After selecting the scattering amplitude of
the circled region and translating it to the origin of the reciprocal space, the
inverse Fourier transform is carried out to get a reconstructed phase image.
When picking up the third term, the center of the sidebands to be shifted to the
origin of the reciprocal space should be carefully selected. The digital
diffractogram of the hologram without specimens can be utilized to find the
center of the sidebands accurately. In the reconstructed phase image
represented by coscp(x, y) in Fig. 2. 23b, the density and the direction of the
52 Daisuke Shindo, Young-Gil Park

white lines indicate the density and the direction (arrows) of the lines of
magnetic flux. In the image, dotted lines indicate magnetic domain walls. It is
also noted that not only the lines of magnetic flux inside the specimen but also
those outside the specimen can be clearly seen in the reconstructed phase
image. In this way, from electron holography the detailed magnetization
distribution and the information about the leakage of magnetic field from the
specimen can be obtained. These informations can hardly be obtained from the
Lorentz microscope image (Fig. 2. 23a) .

2.5.3 Application of Electron Holography


In the following, we present the analyses of the magnetic domain structure in
the hard and soft magnetic materials by electron holography with typical
magnetic materials of NdzFe,4B (Croat et aI., 1984) and Fe735Cu,Nb3Si13sBg
(Yoshizawa et al., 1988), respectively. In each case, the magnetic domain
structures revealed by electron holography will be discussed in relation with
the magnetic properties of each magnetic material. The information obtained
with electron holography is also compared with that obtained with Lorentz
microscopy.
2.5.3. 1 Application to Hard Magnets
It is well known that understanding of the magnetic hardening mechanism is
important in ternary Nd-Fe-B hard magnetic materials. In this section, with
overquenched Nd-Fe-B permanent magnets prepared by the melt-spinning
method, we investigate the cause of the magnetic hardening with the increase
of annealing temperature (Park and Shindo, 2002).
The magnetic properties of the overquenched magnets are shown in
Fig. 2.25. Figures 2. 25a, b show the variation in coercive force as a function
of annealing temperature and the second quadrant demagnetization curves for
three specimens indicated by the circles of A, Band C in Fig. 2. 25a,
respectively. It is seen that the coercive force increases with the increase of
annealing temperature in Fig. 2. 25a. As shown in Fig. 2. 25b, the
demagnetization curve for an as-quenched magnet (A) having a low coercive
force of 378 kA/m decreases rapidly with the applied field. The curve for a
magnet (B) annealed at 843 K having relatively a low coercive force of
621 kA/ m shows a shoulder as indicated by an arrow. This is considered to
consist of two components; one has low coercive force and the other has
relatively high coercive force. The curve for a magnet (C) annealed at 893 K
having a high coercive force of 732 kA/m does not show a dip in contrast with
the demagnetization curve (B). From electron diffraction patterns and electron
microscope images, it is found that an as-quenched magnet is composed of the
amorphous phase, while a magnet annealed at 843 K consists of a NdzFe,4B
crystalline phase and an amorphous phase. The amorphous phase in the latter
Lorentz Microscopy and Holography Characterization of Magnetic Materials 53

magnet is considered to contribute to the component with the rapid decrease in


the demagnetization curve. On the other hand. it is found that a magnet
annealed at 893 K consists of only a Nd2 Fe14 B crystall ine phase. High-
resolution electron microscopy shows no microstructural differences at the
boundaries in the magnets annealed at 843 and 893 K.
, - - - - - - - - - - - - , 1.2

700

! 600
0.8

t:
Cl:l
~ 500
0.4

400
{} -'-------'-----'------'----'
as. quench843 863 883 903 -800 -400 o
Annealing temperature (K) H (kA/m)
(a) (b)

Figure 2.25 (a) Variation in coercive force of an overquenched Ndz Fe!, B magnet as a
function of annealing temperature; (b) Second quadrant demagnetization curves for the
three specimens indicated by the circles of A, Band C in (a), respectively.

Figure 2. 26a shows a Lorentz microscope image of the as-quenched


magnet in an amorphous state. The Lorentz microscope image was observed
by the Foucault mode. and the white and black bands correspond to the
magnetic domains whose sizes are in the range of 200 to 500 nm. By the
Fourier transform. the reconstructed phase image of Fig. 2. 26b can be
obtained from an electron hologram. In the reconstructed phase image of
Fig. 2.26b. it is seen that the direction of the lines of magnetic flux indicated
by arrows slowly changes .in wide region. It is reasonably considered that the
gradual change in the direction of the lines of magnetic flux in the as-quenched
magnet without strong pinning sites corresponds well to lower coercive force
than the annealed ones indicated below.
Figure 2.27a shows a Lorentz microscope image of the magnet annealed
at 843 K. while Fig. 2. 27b shows an enlarged Lorentz microscope image
observed near the specimen edge. In the Lorentz microscop'e image (a)
observed by the Fresnel mode. the bright lines such as those indicated by
arrowheads correspond to the magnetic domain wall contrast. which seems to
be pinned at the grain boundaries. In Fig. 2. 27b obtained by the Foucault
mode. there are many dark and bright grains. It is seen that some of the dark
and bright grains are align along the horizontal line as indicated by arrows of 0
and B. respectively. forming large magnetic domains. In the reconstructed
phase image in Fig. 2.27c. the direction of the lines of magnetic flux gently
changes from place to place especially at the grain boundaries indicated by the
broken lines. It is considered that due to the poor crystallinity at grain
54 Daisuke Shindo, Young-Gil Park

(a)

(b)

Figure 2. 26 (a) Lorentz microscope image of an as-quenched Nd, Fe" B magnet observed
by the Foucault mode; (b) Reconstructed phase image obtained by the Fourier transform.
The upper left part is a vacuum region.

boundaries, the magnetocrystalline anisotropy at the boundaries in the magnet


annealed 843 K is weakened. The gradual change in magnetization distribution
is rather consistent with the magnetic domain structure observed in the Lorentz
microscope image of Fig.2.27b.
Figure 2. 28a shows a Lorentz microscope image obtained by the Fresnel
mode in the magnet annealed at 893 K, while Fig. 2. 28b shows an enlarged
Lorentz microscope image observed near the specimen edge. It is seen that
the magnetic domain wall contrast (arrows) in the magnet annealed at 893 K is
sharper than that in the magnet annealed at 843 K in Fig. 2. 27 a. In the
enlarged Lorentz microscope image of Fig. 2. 28b, magnetic domain walls are
found inside some large grains indicated as "OW." Figure 2. 28c shows a
reconstructed phase image of the same area as that of Fig. 2. 28b. The
reconstructed phase image reveals that the direction of the lines of magnetic
flux abruptly changes at the grain boundaries, being different from the
reconstructed phase image of Fig. 2.27c. Here, we investigate the change of
Lorentz Microscopy and Holography Characterization of Magnetic Materials 55

100nm

(a)

(b)
40011111

Figure 2. 27 Lorentz microscope images of a Ndz Fe'4 B magnet annealed at 843 K


observed by the Fresnel mode (a) and the Foucault mode (b). (c) Reconstructed phase
image.

the direction of the lines of magnetic flux around the grain boundaries in the
reconstructed phase images of Figs. 2. 27c and 2. 28c. Around the grain
boundaries such as indicated by the rectangular regions in Figs. 2. 27c and
2.28c, the angle (e) between two straight lines which are extrapolated along
the lines of magnetic flux in the adjacent grains were measured as shown in the
inset of Fig. 2.29a, b. In the diagrams of Fig. 2.29a, b, the direction and
magnitude of the arrow indicate the deflection angle (e) and the number of the
deflection angles observed in the lines of magnetic flux at the grain boundaries
in Figs. 2. 27c and 2. 28c. While big arrows in the diagram of Fig. 2. 29a are
confined at the small angle less than 20, many big arrows exist in the range
from 30 to 70 in the diagram of Fig. 2. 29b. From this statistical
investigation, it is considered that the magnetocrystalline anisotropy is
enhanced due to the improvement of crystallinity at the boundaries with the
increase of annealing temperature. Although the microstructural change with
the increase of annealing temperature in the two magnets is not identified by
56 Daisuke Shindo, Young-Gil Park

(a)

400 om

OW GB

(b)

(c)

Figure 2. 28 (a), (b) Lorentz microscope images of a Nd2 Fe" B magnet annealed at
893 K observed by the Fresnel mode; (c) Reconstructed phase image. Grain boundaries
and magnetic domain boundaries noted "GB" and "DW" respectively in (b) are indicated
by dotted lines in (c).

using high-resolution electron microscopy, the enhancement of the


magnetocrystalline anisotropy in each grain is clearly identified by electron
holography study.
In addition, the magnetic flux density can be evaluated if thickness of the
crystal is known, since the density of white lines in the reconstructed phase
image directly corresponds to the magnetic flux density. By using electron
energy-loss spectroscopy (Egerton, 1996) assuming the mean free path for
inelastic electron scattering is the same as that for ex-Fe (Zhu and McCartney,
Lorentz Microscopy and Holography Characterization of Magnetic Materials 57

90' 90'

'~
I

~:
-------------_~~_---
180' e o 180' e o
(a) (b)

Figure 2.29 Diagrams showing the deflection angles (e) and their numbers measured at
the grain boundaries in the Nd2 Fe14 B magnet annealed at 843 K (a) and 893 K (b). The
insets in (a) and (b) correspond to the rectangular regions in Figs. 2. 27c and 2. 28c,
respectively. The dotted lines in the insets indicate grain boundaries.

1998; Park and Shindo, 2002), the magnetic flux density of the Nd2 Fe14 8
crystal was evaluated to be about 1. 5 T, being consistent with the saturation
magnetization of the Nd2 Fe14 8 phase, i. e., 1.6 T.

2.5.3. 2 Application to Soft Magnets


In this section, soft magnetic alloys Fe735 CUI Nb 3Si 135 8 gprepared by the melt-
spinning method are used for observing the magnetic domain structures.
Through electron holography study, the change of magnetic domain structure in
thin film specimens with the increase of annealing temperature is investigated.
In addition, by tilting the specimen and introducing the magnetic field into the
thin films, an in-situ experiment is carried out. The results obtained by
electron holography are discussed in terms of the magnetic properties of
Fe735CulNb3Si1358g soft magnetic alloys (Shindo et al., 2002).
Figure 2. 30a, b, c show the coercive force (He)' the permeability (J.lr)
and the magnetostriction (As) of Fe735CulNb3Si1358g alloys with various heat
treatments. The lowest coercive force is obtained at the annealing
temperature of 823 K. It is also noted that the permeability and the
magnetostriction take large and small values, respectively, at this
temperature. While an as-quenched specimen is in an amorphous state,
magnets annealed at 823 K and 973 K consist of fine grains of about 10 nm and
enlarged grains of 50 - 200 nm, respectively. Taking into account the grain
sizes of the specimens annealed at various temperatures, it is noted that the
grain size basically increases with the increase of annealing temperature, but
the size distribution is more broadened at higher annealing temperature.
Figure 2.31 a - d show Lorentz microscope images of Fe735 CUI Nb3Si 135 8 9
films with various heat treatments. The Lorentz microscope images were
observed with the Fresnel mode. In Fig. 2. 31 a of an as-quenched specimen,
rather periodic domain walls with white and black lines are noted. In
58 Daisuke Shindo, Young-Gil Park

103

"I
102
E

'-' 10
~

0
650 700 750 800 850 900 950 1000
As-quenched
Annealing temperature (K)
(a)
14
12
10
~

b 8
x'-' 6
~
4
2
0
650 700 750 800 850 900 950 1000
As-quenched
Annealing temperature (K)
(b)
25

20

15
f'
0 10
x
'-'
5
~
0

-5
650 700 750 800 850 900 950 1000
As-quenched
Annealing temperature (K)
(c)

Figure 2.30 Coercive force (a), permeability (b) and magnetostriction (c) of
Fe73.SCU1Nb3SiI35B9 as a function of annealing temperature.
Lorentz Microscopy and Holography Characterization of Magnetic Materials 59

Fig. 2. 31 b of the specimen annealed at 823 K having the low coercive force
and the high permeability, black and white lines showing the domain walls
change to the broad lines as compared with the Lorentz microscope image of
Fig. 2.31a. With the increase of annealing temperature, the white and black
lines corresponding to the domain walls deviate from the straight-line shape,
but they fluctuate from place to place as shown in Fig. 2. 31 c, d, which
correspond to specimens annealed at 923 and 973 K, respectively.
50011111 50011111

(a) (b)
50011111 50011111

(d)

Figure 2.31 Lorentz microscope images of Fe73.s Cu, Nb 3 Si 13s 8 9 films with different heat
treatments, i. e., as-quenched Ca), annealed at 823 K Cb), 923 K Cc) and 973 K Cd) .

Figure 2.32a-d show the reconstructed phase images of Fe73sCulNb3Si13s89


films, which are obtained from electron holograms with the Fourier transform.
The areas and heat treatments of the specimens in Fig. 2. 32a - d are the
same as those of Fig. 2. 31 a - d, respectively. In Fig. 2. 32a of the as-
quenched specimen, smooth closure domains are clearly seen through the lines
of magnetic flux as indicated by arrows. It is noted that the lines of magnetic
flux are parallel to the specimen edges, thereby eliminating the surface
magnetic charge. Also, the domain walls observed in Fig. 2. 31a just
correspond to the boundaries in Fig. 2. 32d where the directions of the lines of
magnetic flux change at about 90. It is noted that the specimen thickness
gradually increases from the edge to the middle of the specimen, and the
60 Daisuke Shindo, Young-Gil Park

500 nl11

(a) (b)
500 nl11

(c) (d)

Figure 2.32 Reconstructed phase images of Fen 5 Cu, Nb 3 Sin 589 films with different heat
treatment, i. e., as-quenched (a), annealed at 823 K (b), 923 K (c) and 973 K (d). The
white arrows indicate the direction of the lines of magnetic flux. Thicknesses at the positions
indicated by A and 8 with" * ", are estimated from the spacing of white lines.

spacing between the lines of magnetic flux become shorter with the increase of
specimen thickness. If the magnetic flux density of a bulk specimen, i. e.,
1.28 T is assumed, crystal thickness can be simply estimated from
Eq. (2.25). Actually, the thicknesses at the positions of A and B were
estimated to be 36 and 54 nm, respectively. In the above evaluation, we
neglected the phase shift of the incident electrons due to the inner potential, or
in other words we assumed that the crystal thickness does not change
appreciably at the regions A and B. In Fig. 2. 32b of the specimen annealed at
823 K, the lines of magnetic flux curve gradually forming rather circular
magnetic domains, being different from Fig. 2. 32a, where the lines of
magnetic flux take a rather straight-I ine shape. It is also noted that some of the
lines of magnetic flux do not close and there is some leakage of the lines of
magnetic flux outside the specimen being seen especially at the bottom of the
figure. The reason of the leakage of the lines of magnetic flux is considered as
follows. Although the residual magnetic field is basically perpendicular to the
specimen, there exists some weak external magnetic field along the film plane
Lorentz Microscopy and Holography Characterization of Magnetic Materials 61

due to the wedge shape and surface roughness of the specimen. Since the
external magnetic field is the same and the thickness and morphology of the
specimens are considered to be similar in the present study, the leakage of the
lines of magnetic flux in the specimen annealed at 823 K is considered to result
from its higher permeability than that of the as-quenched specimen. On the
other hand, in Fig. 2.32c, d of the specimens annealed at higher
temperatures, sharp differences from Fig. 2. 32c, b are noted. The size of
magnetic domains becomes smaller, and the lines of magnetic flux deviate
significantly from the smooth line shape. As reported previously (Yoshizawa
and Yamauchi, 1991), by annealing at high temperature, in addition to the
crystalline phase of bcc Fe containing Si and B, various Fe-B compounds form
and the size of these grains increases gradually with the increase of annealing
temperature. Thus, the irregularity in the shape of the lines of magnetic flux is
considered to result from the inhomogeneous magnetization distribution due to
the bcc Fe and the Fe-B compounds.
Figure 2. 33 shows the reconstructed phase images of the Fen 5 CU, Nb3 Sin 5 B9
film annealed at 823 K with the increase of tilting angle for the specimen. Due
to the ti Iting of the specimen, the component of the external magnetic field
along the film plane will increase gradually. As indicated in the figures, with
the increase of the tilting angle, the lines of magnetic flux forming the closure
domain change gradually and eventually tend to be almost parallel due to the
0
effect of the external magnetic field at the titl ing angle 4 It is also noted that

the leakage of the lines of magnetic flux increases with the increase of the
tilting angle.

Figure 2.33 Reconstructed phase images of an Fe73SCu, Nb3Si,3SBg film annealed at


823 K with the increase of tilting angle. In-plane components of the external magnetic field
are 0 Aim (tilt 0), 2.8 Aim (tilt n, 8 3 Aim (tilt 3), 11.1 Aim (tilt n.
62 Daisuke Shindo, Young-Gil Park

Figure 2. 34a, b show the reconstructed phase images of the as-quenched


Fen 5 CUI Nb3Si 13 5 B9 film for the increase of tilting angle. In the tilting angle of
3, the lines of magnetic flux are strongly affected by the tilting. The flow of
the lines of magnetic flux changes to a rather circular shape due to the tilting.
It is interesting to note that the magnetic domain structure in the tilted
specimen is similar to that of the specimens annealed at 823 K with no tilt
shown in Fig. 2. 32b. Thus, it is confirmed that the effect of the external magnetic
field is smaller in the magnetic domain structure of the as-quenched specimen than
that of the specimen annealed at 823 K. This corresponds well to the relatively large
coercive force and the small permeability of the as-quenched specimen compared
with the specimen annealed at 823 K as shown in Fig. 2. 30a, b. Figure 2. 34c, d
are the reconstructed phase images of the F~3 5 CUI Nb3Si 13 5 B9 film annealed at 973
K for the increase in tilting angle of the specimen. Being different from Figs. 2. 33
and 2. 34a, the lines of magnetic flux do not change so much. The difference directly
indicates the strong pinning of magnetic domain walls due to the precipitates of
Fe-B compounds at the grain boundaries of the bcc Fe phase, resulting in the
drastic increase of the coercive force and the decrease of the permeability.

TiitY

(a)

(c)

Figure 2. 34 ( a ), (b) Reconstructed phase images of an as-quenched


Fe735Cu,Nb3Si135Bg film for the increase of tilting angle; (c), (d) Reconstructed phase
images of a Fe735CulNb3 Si'35B9 film annealed at 973 K for the increase of tilting angle;
In-plane components of the external magnetic field are 8.3 Aim (b) and 16.6 Aim (d).
Lorentz Microscopy and Holography Characterization of Magnetic Materials 63

2.6 Concluding Remarks

In addition to analytical electron microscopy, we have explained Lorentz


microscopy and electron holography which are quite useful to analyze the
complicated magnetic domain structure and thus to understand the magnetic
properties of advanced magnetic materials. Since not only cold FEGs but also
thermal FEGs are utilized for TEM observation with high brightness and high
coherency, currently it is not so difficult to obtain detailed magnetic domain
structure especially by DPC Lorentz STEM and electron holography. It is also
pointed out that peripheral instruments such as electron spectrometers and
recording systems are highly developed, and thus thickness evaluation and
image contrast analysis can be accurately carried out. The quantitative
analysis of magnetic domain structure will be expected accordingly. The
authors believe that magnetic domain structure in nanometer scale will be
clarified quantitatively in the near future.

References
Aharonov, Y. and D. Bohm, Phys. Rev. 115: 485 (1959)
Allen, S. M. Phil. Mag. A 43:325 (1981)
Boersch, H. and H. Raith. Naturwissenschaften 46: 574 (1959)
Chapman, J. N. and E. H. Darlington. J. Phys. E 7: 181 (1974)
Chapman, J. N., P. E. Batson, E. M. Waddell and R. P. Ferrier.
Ultramicroscopy 3 :203 (1978)
Chapman, J. N., A. B.Johnston, L. J. Heyderman, S. McVitie, W. A. P.
Nicholson and B. J. M. Bormans.IEEE Trans. Magn. 30:4479 (1994)
Cohen, M. S. J. Appl. Phys. 38 :4966 (1967)
Croat, J. J., J. F. Herbst, R. W. Lee and F. E. Pinkerton. J. Appl. Phys.
55: 2078 (1984)
Egerton, R. F. Electron Energy-Loss Spectroscopy in the Electron
Microscope, 2nd edn. Plenum Press, New Yorke 1996)
Endo, J. T. Matsuda and A. Tonomura. Jpn. J. Appl. Phys. 18:2291 (1979)
Frost, B. G., L. F. Allard, E. V61kl and D. C. Joy. In: A. Tonomura, L. F.
Allard, G. Pozzi, D. C. Joy and Y. A. Ono Electron Holography, Elsevier
Science B. V., p169(1995)
Fuller, H. W.and M. E. Hale.J. Appl. Phys. 31:238 (1960)
Gabor, D.Proc. Roy. Soc. (London) A 197:454 (1949)
Hale, M. E., H. W. Fuller and H. Rubinsten. J. Appl. Phys. 30:789 (1959)
Hirsch, P., A. Howie, R. Nicholson, D. W. Pashley and M. J. Whelan.
Electron Microscopy of Thin Crystals. Krieger Publishing Company( 1977)
64 Daisuke Shindo, Young-Gil Park

Huber, E. E. Jr. D.O. Smith and J. B. Goodenough. J. Appl. Phys. 29: 294
(1958)
Joy, D. C., Y. -So Zhang, X. Zhang, T. Hashimoto, R. D. Bunn, L. Allard
and T. A. Nolan. Ultramicroscopy 51: 1 (1993)
Kappert, H., Z. Angew. Phys. 37: 139 (1970)
Kappert, H. and P. Schmiesing. Phys. Stat. Sol. 4: 737 (1971)
Kelly, P., M.A. Jostsons, R. G. Blake and J. G. Napier. Phys. Stat. Sol.
(a) 31:771 (1975)
Kittel, C.Rev. Mod. Phys. 21:541 (1949)
Lorimer, G., G. CI iff and J. N. Clark. Development in Electron Microscopy
and Analysis. Acad. Press, London, p. 153( 1976)
Matsumoto, K and M. Takahashi. J. Opt. Soc. Am. 60: 30 (1970)
Mayer, L.J. Appl. Phys. 28:975 (1957)
McCartney, M. R. and Y. Zhu. Appl. Phys. Lett. 72: 1380 (1998)
McVitie, S. J., N. Chapman, L. Zhou, L. J. Heyderman and W. A. P
Nicholson.J. Magn. Magn. Mater. 148:232 (1995)
Mishra, R., K.G. Thomas, T. Yoneyama, A. Fukuno and T. Ojima.J. Appl.
Phys. 52:2517 (1981)
Mooney, P. E., G. Y. Fan, C. E. Mayer, K. V. Truong, D. B. Bui and O.
L. Krivanek. Proc. of the 12 th International Congress for Electron
Microscopy. San Francisco Press, Seattle 1: p. 164 (1990)
Morrison, G. R., J. N. Chapman and A. J. Craven. Inst. Phys. Cont. Ser.
No. 52 257 (1980)
Nagel, H. J. Appl. Phys. 50: 1026 (1979)
Neel, L.Compt. Rend. 241: 533(1955)
Park, Y.-G., D. Shindo and M. Okada. Materials Transactions, JIM. 41:1132
(2000)
Park, Y. -G., D. Shindo, H. Kanekiyo and S. Hirosawa. Materials
Transactions, JIM. 42: 1878 (2001)
Park, Y. -G. and D. Shindo. J. Magn. Magn. Mater. 238: 68 (2002)
Perkins, R. S., S. Gaiffi and A. Menth, IEEE Trans. Magn. MAG-ll:1431
(1975)
Pozzi, G.andG. F. Missiroli.J. Micro. 18:103(1973)
Reimer, L.and H. Kappert. Z. Ang. Phys. 26:58 (1969)
Reimer, L. Transmission Electron Microscopy. Springer-Verlag (1993)
Rogulic, M. Ph. D. Thesis. Cambridge University (1964)
Shindo, D. and K. Hiraga. High-Resolution Electron Microscopy for Materials
Science. Kyoritsu Press, Tokyo, p. 138( 1998)
Shindo, D., Y. Ikematsu and Y. Murakami. JEOL news 35 E: 10 (2000)
Shindo. D., Y. -G. Park and Y. Aoyama. Int. Cont. on Composites
Engineering, Spain (2001) p. 847
Shindo, D., Y.-G. Park andY. Yoshizawa.J. Magn. Magn. Mater. 238:101
(2002)
Sonoda, M., M. Takano, J. Miyahara and Y. Harada. Radiology 148: 833
Lorentz Microscopy and Holography Characterization of Magnetic Materials 65

( 1983)
Stenton, N. M., R. Notis, J. I. Goldstein and D. B. Williams. Quantitative
Microanalysis with High Spatial Resolution. The Metals Soc. London, p.35
( 1981)
Suzuki, T. and A. Hubert. Phys. Stat. Sol. 38:K5 (1970)
Suzuki, T, K. Hiraga and M. Sagawa.Jpn. J. Appl. Phys. 23:L421 (1984)
Taniyama, A.D. Shindo and T. Oikawa. J. Electron Microsc. 46:303 (1997)
Tanji, T, K. Urata, K. Ishizuka, Q. Ru and A. Tonomura. Ultramicroscopy
49:259 (1993)
Tonomura, A. Jpn. J. Appl. Phys. 11 :493 (1972)
Tonomura, A, T. Matsuda, J. Endo, T. Arii and K. Mihama. Phys. Rev.
Lett. 44: 1430 (1980)
Tonomura, A.Adv. Phys. 41:59 (1992)
Tonomura, A. Electron Holography, 2nd Edn. Springer-Verlag (1999)
Tsuno, K. and M. Inoue, Optik 67: 363 (1984)
Tsuno, K.Rev. Solid State Sci. 2:623 (1988)
Tsuno, K, M. Inoue and K. Ueno, Mater. Sci. Eng. B 3:403 (1989)
Wade, R. H.J. Appl. Phys. 37:366 (1966)
Yang, J.-M., D. Shindo and H. Hiroyoshi. Materials Transactions, JIM. 38:
363 (1997)
Yang, J. -M., D. Shindo, S. -H. Lim, M. Takeguchi and T. Oikawa. Electron
Microsc., ICEM 14:559 (1998)
Yoshizawa, Y., S. Oguma and K. Yamauchi. J. Appl. Phys. 64: 6044 (1988)
Yoshizawa, Y. and K. Yamauchi. Materials Transactions, JIM. 31: 307 (1991)
Zhu, Y. and M. R. McCartney. J. Appl. Phys. 84: 3267 (1998)

The authors wish to thank Dr. Y. Yoshizawa, Hitachi Metals, Ltd. and Dr. K. Tsuno, JEOL
Ltd. for providing the samples of Fen5 CUI Nb3 Si 135 8 9 and the data of DPC Lorentz STEM,
respectively. They also thank Dr. J. -M. Yang for his cooperation in Lorentz microscopy and
electron holography.
3 Characterization of Magnetic Materials by
Means of Neutron Scattering

G. Ehlers, F. Klose

3. 1 Introduction

The neutron is an elementary particle with a particular combination of


properties that makes it uniquely versatile as a condensed matter probe.
These properties are:
(1) It is not electrically charged;
(2) It has spin 1/2 and a magnetic dipole moment of JJ~ = -1. 913JJN(with
the nuclear magnetonJJN = 5.051 x 10 -27 J/T) ;
(3) It has mass (m n = 1.675 X 10- 27 kg).
The lack of an electric charge has two important consequences. First,
since the dominant scattering by electrical charges is absent, in solid matter
one can observe neutron scattering processes by atomic nuclei via strong
interaction or by magnetic moments via magnetic interaction with the magnetic
moment of the neutron. One of the most remarkable features of the neutron
scattering technique is that magnetic interaction strengths are often on the
same order of magnitude as nuclear interactions such that magnetic and
structural properties of the sample can be probed simultaneously. Second, the
neutron is able to penetrate deeply into condensed matter (typically a few mm
to a few cm) unless there is one of the few isotopes with high absorption cross-
section present. The neutron is therefore well suited for the study of true bulk
properties of matter.
Magnetic scattering experiments can be performed on a large variety of
samples, including, for example, single crystals, powders, and even
artificially grown thin film structures. Physical quantities that can be measured
by magnetic neutron scattering include (see Sections 3. 2 and 3. 3 for a
detailed discussion) :
( 1) Magnetic structures across various length scales (e. g., ranging from
ferro-and antiferromagnetic order at atomic length scales to mesoscopic
structures like flux lattices of superconductors or magnetic depth profiles in thin
films) ;
(2) Magnetization density (e. g., how the moment is distributed in the
vicinity of magnetic atoms) ;
Characterization of Magnetic Materials by Means of Neutron Scattering 67

(3) Magnetic excitations (e. g . spin waves).


In many cases the obtained information is unique. i. e . it cannot be
derived by means of other experimental techniques.
Neutron beams are produced in research reactors by nuclear fission (for
example at the Institue Laue-Langevin (ILL). Grenoble. France. see
www.ill.fr) or in spallation sources by using linear proton accelerators (for
example at ISIS at the Rutherford Appleton Lab. Oxford. Great Britain. see
www.isis.rl.ac. uk. or at the US Spallation Neutron Source (SNS). see
www.sns.anl. gov ) . Although neutron scattering is an intensity-limited
technique (see Section 3. 1. 1) one can obtain sufficient intensity (at today's
best sources typically 10 6 - 10 7 neutrons cm- 2 S-1 on the sample) to do
experiments. The neutron has mass and this has the important consequence
that neutrons created by these processes can be efficiently slowed down
(moderated) by collisions with light atoms to specific energy ranges that are
favorable for particular condensed matter studies. The energy ranges are
typically labeled according to the temperature of the moderator material.
Heated graphite blocks (T > 2000 K) deliver "hot" neutrons. room-
temperature water moderators deliver "thermal" neutrons and liquid hydrogen
moderators (T <30 K) are used for "cold" neutron production. The relation
between the neutron kinetic energy and (via de-Broglie) its wavelength is
p2 h2
E = - = - - 2.
2m 2mA
81.8042
E (meV) = [A (A) J2 .

Thus. typical ranges for neutron energies and corresponding wavelengths used
in neutron scattering experiments are (approx. ) :
(1) "hot" neutrons E is 100-500 meV. A is 0.5-1 A;
(2) "thermal" neutrons E is 10- 100 meV. A is 1- 3 A;
(3) "cold" neutrons E is 0.1-10 meV. A is 3-30A.
We see that both energy and wavelength are in ranges that are suitable
for condensed matter investigations. the wavelength being in the range of
typical atomic distances and the energy corresponding to that of lattice
vibrations or spin excitations. There is thus a fundamental difference between
neutrons and X-rays. where a wavelength of 1 A corresponds to the energy of
12.4 keV. Neutrons can therefore probe correlations in condensed matter in
space (atom ic to mesoscopic length scales) and time (10 - 14 s to 10 -7 s)
simultaneously. Figure 3. 1 illustrates some important characteristic length
scales for structural and magnetic properties of condensed matter. In
scattering experiments. the connection between probed length scale d and
neutron wavelength A is given by Bragg's law: 2 d sine = n A. where 2e is
the scattering angle and n is an integer.
The neutron as a free (i. e . unbound) particle is not stable. Currently.
the best value for neutron Iife time is (886. 7 1. 9) s (Groom et al.. 2000).
68 G. Ehlers, F. Klose

Exchanue Magnetic dipolar


I", .'" "I
I Screel1lng .. I I .interactior~ I
RKKY I ~pin ditrusio~11
I"

Monolayer Layer thickness


1......0 - - - - - - - - - ; : : - ' - - - - - ; - - - - - ; - - - - ; - : - - - ; - - - - - - - -
Supertattice periodicity
Atomic 1.... 0 - - - - - - - - ' - - - - - ' - - - - - - - - - ' ' - - - - - - - - -
diameter lattice constants
r-I" "I

Neutron wavelength
I"
Neutron diffraction .. I Neutron reflectolllctry
I"

Ih, I,
lOA
1

100A
0

lOOOA
I

Figure 3. 1 Comparison of characteristic length scales of magnetic and structural


properties of matter and length scales that are accessible in typical neutron scattering
experiments.

This is, in practice, however, not an issue because the life time is much
longer than necessary to do scattering experiments, i. e., the time the neutron
spends travel ing from the source to the detector.
The intention of our article is to give an overview of basic concepts of
magnetic neutron scattering techniques at an introductory level. We will
emphasize the experimental aspects and will discuss some recent scientific
highlights that illustrate the uniqueness of these techniques. We do not intend,
however, to review the scientific progress in this large research field. Our
choice is a personal one and the available space is limited, so we will
certainly omit some topics which might be of interest to a particular reader.
For more details and other aspects of neutron scattering, we refer to
introductory texts available in the literature (de Gennes, 1963; Rossat-
Mignod, 1987; Mezei, 1991), or to one of the numerous text books on neutron
scattering (Squires, 1978; Lovesey, 1984; Williams, 1988; Furrer, 1995).
Our article is organized as follows: In the first section we will briefly
discuss the basic magnetic scattering theory (Section 3. 1. 1) and give an
introduction to polarized neutron beam instrumentation (Section 3. 1.2) as well
as to polarization analysis (Section 3. 1. 3). Section 3. 2 and 3. 3 review
neutron scattering instruments that are particularly useful for studies on
magnetic materials. Section 3.2 focuses on elastic magnetic scattering and in
Section 3. 3 we discuss spectrometers that are specifically designed to
measure excitations in magnetic materials. Scientific examples are included
for both types of instruments.
Characterization of Magnetic Materials by Means of Neutron Scattering 69

3. 1. 1 Cross Section Formalism


Nuclei, with a typical diameter on the order of femtometers (10 -15 m), are
point-like on the scale of typical neutron wavelengths in the A range (10- 10 m) .
Therefore the scattering from a fixed nucleus is essentially s-wave scattering
and can be described by a single (generally complex) number b, the
scattering length. This parameter is known for most isotopes to a precision of
a few percent and is on the order of 10 - 13 to 10 - 12 cm (a Iist of neutron
scattering length and cross sections can be found here: http://www . ncm.
nist. gov/resources/n-Iengths/). In comparison, typical b values for X-rays
are on the order of 10 - 11 cm, resulting in much stronger interactions in
scattering experiments. In contrast to the X-ray case, however, there is no
systematic dependence of b on the atomic number Z and nucleon number N.
This opens the possibility to study isotope effects. The case of hydrogen and
deuterium is of particular importance for soft matter physics. Isotope mixing
can also be beneficial in magnetism studies, for example for enhancing the
sensitivity for magnetic signals (Hoffmann et aI., 2002).
From these numbers one can immediately conclude that typical cross
sections a = 41Tb 2 are weak and that, therefore, the scattering process may
be treated in Born approximation of perturbation theory. This picture breaks
down only for a few exceptional cases, for example coherent Bragg scattering
from perfect crystallites of dimension d::::::::10- 3 cm or larger (Mezei, 1991) or
grazing incidence techniques like neutron reflectometry (see Section 3.2.3).
Due to the relatively small values for the scattering lengths and t~e
correspondingly weak beam-sample interactions, neutron scattering is an
intensity-limited technique. This has its consequences in the instrument
design, where one often tries to increase intensity by using large detector
arrays covering a large solid angle. On the other hand, a convenient corollary
is that data are relatively easy to interpret because multiple scattering
contributions are usually small.
Magnetic scattering results from the interaction of the neutron spin with
the magnetic field generated by the unpaired electrons of the atom. The
corresponding magnetization density function may be strongly delocalized, its
extent often being comparable in size with typical neutron wavelengths.
Therefore, in contrast to nuclear scattering, scattering from a magnetic
moment depends on the momentum transfer, resulting in a magnetic form factor
F ( q) which suppresses intensities towards high q (see below).
Nevertheless, one can define an effective magnetic scattering length p as
well, which, for a magnetic moment of J.l B' amounts to 0.27 x 10 -12 cm.
In a neutron scattering experiment one measures a cross section, that is,
the number of neutrons scattered per unit time into a solid angle dO with
neutron energy transfer in the interval [itw, it (w + dw) J, normalized to the
70 G. Ehlers, F. Klose

incident neutron flux. Figure 3.2 shows a typical scattering geometry.

Sample

Figure 3.2 Scattering geometry: The incident neutrons are traveling along the z axis and
I
are characterized by their wave vector k i and energy Ei ( I k i = ~2mEi In ). After
interaction with the sample, a fraction of the incident beam is scattered into a small solid
angle dO (the angles 28 and (1) define the direction of the scattered beam). As a result,
the final wavevector is k f and the energy E,. q is the corresponding wave vector transfer.
The number of scattered neutrons may be monitored by a detector at a distance r that
covers this range of sol id angle.

The above mentioned cross section is given by Fermi's Golden rule,

where v and v' are initial and final states of the scatterer with corresponding
energies E v and E v" m n is the neutron mass, k i and k f are initial and final
neutron wave numbers, and p v is the population factor of the initial states.
The interaction potential II is the sum of two parts: one arising from nuclear
interaction with the nuclei

2n n 2 " " '


II nue, = ---LibioCr - rJ
mn i

which we are not interested in here, and a second arising from the interaction
with the magnetic field B in the sample

IImagn=-ynsoB
s
where is the operator of the neutron spin and y /2n = - 2916.4 Hz/G is the
gyromagnetic ratio of the neutron. The matrix elements can be evaluated
CLovesey, 1984; Williams, 1988) and with some assumptions Cunpolarized
beam, no orbital moments contribute to B) one arrives at
Characterization of Magnetic Materials by Means of Neutron Scattering 71

(the case of scattering from orbital moments is more complicated but does not
contain new features). This expression contains the momentum transfer q, the
classical electron radius ro = O. 282 X 10- 12 cm , the magnetic form factor F (q)
(see below), the Debye-Waller factor e- 2W(q) , and the magnetic scattering
function
""
S"~(q, w) = -l-fe-iwtdt~ejq(r,-r')<Sf(O)Sf(t
2TT h -00
..
11

where Sf( ex = x, y, z) is the spin operator of the ion at site rj and brackets
mean the thermal average. Some important features can be read from these
formulae. First, and most important, the factor

( o,,~ - q;~~)

implies that only components of the magnetic moment density perpendicular to


q are relevant in magnetic scattering. Second, the form factor F (q), which is
the Fourier transform of the spin density associated with the magnetic moment,
is monotonically falling off with q and suppresses magnetic scattering intensity
towards high q. Form factors can be measured using polarized neutrons
(Brown et aI., 1999) or calculated (Desclaux and Freeman, 1984) and can be
found in tables (Brown, 1992). Last but not least, S"~(q, w) is the Fourier
transform in space and time of a pair correlation function giving the probability
that if the magnetic moment at site rj has some vector value at time 0, then the
moment at site rj has some other value at time t.
A similar interpretation of the scattering cross section in terms of
correlation functions is also valid for nuclear scattering.

3.1.2 Polarized Neutron Beam Instrumentation


This chapter describes the basic means used to polarize a neutron beam and to
handle the polarization. A neutron beam is said to be polarized if the beam
average of the individual polarization vectors of all neutrons is non-zero. The
beam polarization P can be determined by measuring the numbers of spin "up"
( N up) and spin" down" (N down) neutrons in the beam:

P = N up - N down .
N up +N down

The common methods to polarize neutrons are:


(1) Transmission or reflection from magnetic mirrors or supermirrors;
(2) Bragg diffraction from a magnetic crystal;
72 G. Ehlers, F. Klose

(3) Transmission through a polarized 3 He filter.


We shall describe these methods briefly. More detailed reviews can be
found in the literature (Williams, 1988; Cussen et aI., 2000). Let us note here
that neutron sources of the next generation (I ike the SNS under construction at
Oak Ridge National Laboratory) will produce pulsed beams. Here neutrons are
effectively monochromatized by their time-of-flight between moderator,
sample and detector, because the velocity is inversely proportional to the
wavelength. Monochromators are therefore not appropriate for instruments at
a pulsed source (exceptions are, for example, energy analyzing crystals).
However, 3 He spin filters are well suited for pulsed beams and also
supermirror based polarizers can be designed to handle broad bandwidth
neutron beams (Krist et al., 1998).
Magnetic mirror polarizers use total reflection of neutrons from magnetic
surfaces or thin films (Fig. 3. 3). The refractive index of a magnetized
material contains two parts, one due to the nuclear scattering length density
and one due to the magnetic. The latter is proportional to the magnetization M
and depends on the orientation of the neutron spin with respect to the vector
M. Hence it is different for up ( +) and down ( -) neutrons. The critical angle
for total reflection of neutron on a magnetic surface is

where i\ is the neutron wavelength, n the number of atoms per unit volume,
and band p are the average nuclear and magnetic scattering length,
respectively. Thus, for a neutron beam that strikes the surface within the
angular range between e; and e: (note that these angles depend on the
neutron wavelength) , the reflected beam will practically consist only of spin up
neutrons and therefore be highly polarized. By alloying, one can match band
p such that one spin state (in most cases spin down) is not reflected at all and
thus create a polarized reflected beam for an incident angular range between
eO and e:. In the case that a neutron transparent substrate is used, one can
also use the transmitted beam since it is polarized as well (Fig. 3. 3). Such a
device is effectively a "spin splitter" (Krist et al., 1998). Unfortunately,
typical values for band p are very small such that polarizers based on single
magnetic layers-even for large neutron wavelengths of several A-only allow
an angular coverage of a few mrad.
A large improvement was made by artificially increasing the critical angle
by using multilayer Bragg reflections from a "supermirror" structure (Mezei
and Dagleish, 1977). A typical polarizing supermirror consists of a magnetic/
non-magnetic layer sequence in which for spin down neutrons the scattering
length density (SLD) of the magnetic layer matches exactly the SLD of the
non-magnetic layer (a possible combination is Fe89Col1 as the magnetic and Si
as the non-magnetic material (Krist et al., 1988. If this is the case, then
spin down neutrons do not experience any contrast (or potential difference) at
Characterization of Magnetic Materials by Means of Neutron Scattering 73

Incident
lInpolarized beam

Transmitted
spin "down"
polarized beam

Figure 3. 3 Magnetic thin film polarizer. The external magnetic field Be,' defines the
direction of the neutron polarization. Be,' must be strong enough to fully saturate the film
magnetization M.

the interfaces of these materials and are, consequently, not reflected by the
multilayer structure. In contrast, spin up neutrons experience a strong contrast
between the layers such that they are strongly reflected. A supermirror
layering sequence does not consist of only one single multilayer period since
this would only reflect a particular wavelength/angle of incidence combination.
Instead, the multilayer period is slowly varied throughout the layering
sequence. This allows handl ing of a larger range of wavelength/angle
combinations.
Nevertheless, there is a natural limitation of the method that can be seen
in the formula: even for state-of-the-art supermirrors, for thermal neutrons ee
is still in the mrad range (for polarizing coatings, up to three times the critical
angle of natural Ni has been achieved), therefore they work best for cold
neutrons, but the accepted angular range is always relatively small. Modern
mirrors reach typical beam polarizations of 95 % - 98 % at a neutron
transmission of the desired spin state of about 90 % .
The (111) reflection from the ferromagnetic Heusler alloy CU2 MnAI
(Freund et al., 1983) can be used to polarize thermal neutrons. This is
convenient for instruments that use a monochromatic beam (diffractometers
and triple-axis spectrometers at reactor installations for example) because
polarization and monochromatization are done simultaneously. If the crystal is
placed in a vertical field, the magnetic moments are aligned along the field
direction and the magnetic scattering is entirely non-spin-flip. The structure
factor of the (111) reflection gives scattered intensities "up" and "down" of
/+oc(b AI -b Mn +PMn)2 and /-oc(b AI -b Mn -PMn)2, respectively, where the
b 's are the nuclear scattering lengths of the elements and PMn is the magnetic
scattering length. The particular property of this alloy is that b AI - b Mn ~ PMn'
so the spin down intensity is small and a polarization of 95 % can be routinely
achieved. The drawback is that the reflected intensity is still much smaller
than for a pyrolytic graphite monochromator (the loss factors are about 5 at
74 G. Ehlers, F. Klose

2.4 A and about 10 at 1.5 A), mainly because of the smaller mosaic spread of
the available crystals.
For cold neutron instruments it is, in the majority of cases, better to
polarize with a supermirror and monochromatize independently using a
pyrolytic graphite crystal (unless Heusler crystals with sufficiently large
mosaicity are available). This approach has the additional advantage that,
due to the low neutron absorption of graphite, the spectrum transmitted by the
monochromator can be used to feed neutrons to other instruments downstream
the neutron guide. The decoupling of monochromatization and polarization is
particularly important for triple-axis spectrometers since it allows for a more
flexible resolution setting.
Polarizing filters based on preferential scattering or absorption of one of
the neutron spin states have been known for decades (Williams, 1988).
Various techniques have been developed but none of them have become widely
accepted for various reasons (insufficient polarizing efficiency, unfavorable
neutron energy-dependence, low transmittance, too compl icated setup, high
cost, etc.). A more recent and very promising development, however, is the
helium spin filter. Although it has been recognized since the 1960s that
polarized 3He gas might be an extremely useful spin filter for thermal and
epithermal neutron beams (Passell and Schermer, 1966), the feasibility could
be demonstrated only recently (Coulter et aI., 1990). The device is based on
the spin dependence of the neutron absorption by the 3He isotope, At neutron
energies typically used in neutron scattering, absorption is largely dominated
by the reaction 3He t +n t =:>4He* =:>IH+ 3H, which goes entirely through the
singlet state of the compound nucleus with zero spin, that is, only the neutron
spin state with spin antiparallel to the 3He spin contributes to the absorption
cross section for the capture of neutrons. At a neutron energy of 25 meV, the
cross section is 10 666 barns for neutrons with spin anti parallel to the 3He
nuclear spin, while the total cross section for absorption or scattering of
neutrons with parallel spin is only a few barns (Mughabghab et al., 1981).
This makes it possible to construct a spin filter device provided the 3He can be
polarized. For a sufficient column density (atomic density x length of the cell)
of 100 % polarized 3He, the transmission of neutrons with parallel spin would
approach 100 %. This transmitted neutron beam would be 100 % polarized
since virtually no neutrons with anti parallel spin could pass through the cell.
Highly polarized 3He can be obtained by two methods that use optical
pumping techniques: spin exchange with optically pumped Rb, and
metastability exchange optical pumping in 3He. In recent years, both
approaches have been used successfully in polarization and analysis
experiments of neutron beams (Heil et aI., 1999; Cussen et al., 2000; Rich
et al., 2001; references therein) .
3He spin filters show many favorable characteristics: CD they are
appropriate for polarizing/analyzing neutrons of a very broad energy range
from cold to epithermal, (Z) they are suitable for broad wavelength band
Characterization of Magnetic Materials by Means of Neutron Scattering 75

experiments. @ their beam polarization is highly homogeneous. they have


a predictable transmission function. @ a large solid angle of detection is
possible. and @ their additional contribution to the beam divergence is
minimal. Due to these features. the main application of 3He is polarization
analysis in experiments that require large angular coverage. e. g . small-angle
scattering. diffuse scattering. off-specular neutron reflectometry. etc. For
these applications. 3He is clearly easier to use compared to a complicated and
expensive arrangement of supermirrors to cover large solid angles. In
contrast. supermirrors are usually superior in experiments that require only
small angular coverage at lower neutron energies. e. g. specular
reflectometry.
3He spin filters are. however. not yet fully mature devices. Issues that
require further technical development include: CD improving the 3He
polarization reliably beyond the current experimental limit of approximately
60 % - 70 %. <Z) achieving reproducible and very long relaxation times.
@ fabricating large solid angle cells. and shielding the cell from magnetic
field gradients resulting from high sample fields. The first point is of particular
importance. Due to the non-perfect 3He polarization. the "wrong" neutron
polarization state (anti parallel) is partially transmitted. whereas the desired
parallel neutron spin state is partially absorbed. Therefore. in order to
achieve high neutron beam polarization. a relatively large column density of
3He gas is required. which implies low transmission. A state-of-the-art 3He
cell designed for achieving a neutron beam polarization of 95 %. a typical
value for a polarizing supermirror. will provide approximately three times less
neutron flux than a supermirror polarizer.
To manipulate the neutron spins of a polarized beam. a magnetic field is
needed. and the time evolution of an individual spin in such a field can be
classically described by the well-known Larmor equation.

~tS (t) = y[s( t) x Bet) ]

where y/21f= - 2916. 4Hz/G is the gyromagnetic ratio of the neutron. B is


the magnetic field and s is the spin vector of the neutron. B (t) refers to the
time evolution of the magnetic field that the neutron experiences along its
trajectory. i. e . in the stationary frame of the moving neutron. Remembering
that the neutron is a spin 1/2 particle. and conventionally taking the direction
of the B vector as the z axis. the two most important solutions of this equation
are:
(1) sct = 0) II B. in this case the operator of the spin is in a z eigenstate
and hence s (t) is constant. Simple polarization analysis basically uses this
case (with the exception of a necessary spin flipper device).
(2) s ( t = 0) -.l B in this case s (t) is not constant. and the beating
between x and y states results in what is classically understood as Larmor
precession of the spin. The most prominent instrumental application of Larmor
76 G. Ehlers, F. Klose

precession of a neutron beam is neutron spin echo (see Section 3.3.2).


Case (1) means that the polarization, once" prepared" to be parallel to
the field direction, will stay parallel to the field (hence the term guide field for
B), as long as the direction of the B-field does not change. If the field
direction does change, say, in the frame of the moving neutron with a typical
frequency w, two cases have to be considered that relate W to WL = yB, the
Larmor frequency defined by the modulus of the field:
(1) wwL(slow or"adiabatic" field change), as for example in a field
that is spatially changing direction slowly along the neutron trajectory. The
neutron spin follows the field rotation adiabatically and essentially stays
parallel to the local field direction.
(2) WWL' (sudden or "non-adiabatic" field change), as for example if
the neutron beam passes through a thin metal foil that carries a dc current (the
magnetic field vectors before and behind the current sheet point in opposite
directions). In cases when there is a non-zero angle between the direction of
the neutron spin and the new B -field vector, the neutron spin starts precessing
around this new local field direction. This is made use of in spin fl ipper
devices.
Technically, guide fields of a few Gauss are sufficient to prevent the earth
field or stray fields from neighboring installations from disturbing the local field
direction.
Besides a magnetic guide field that is necessary to keep the polarization,
other devices are often needed which turn the polarization direction with
respect to the guide field direction. Generally, this is achieved by making use
of non-adiabatic (sudden) changes of the field direction which the neutron spin
can not follow. Different types of these "spin flippers" are in use. A simple

Precession inside the coil

B""t
c=)-------
t
------,

t
Neutron
beam
Spin

B int
Figure 3.4 Schematic diagram of a Mezei-type spin flipper. B;ot is chosen such that
the polarization is inverted, i.e., turned by "IT, when the neutron exits the flipper. The
additional coil that cancels the guide field within the flip-field region is omitted. The
windings of this coil will be perpendicular to the windings of the flip field.
Characterization of Magnetic Materials by Means of Neutron Scattering 77

example is the Mezei flipper, which is basically a flat coil placed in the beam
which creates a field highly localized in its interior. This flipping field Bini is
perpendicular to the external guide field B ext (Fig. 3. 4). It is important to note
that in this design the external guide field is cancelled within the flip field
region by a second coil placed around the flipper coil. When the neutron enters
the flipper, it experiences a sudden change of the field direction, such that the
spin precesses around the inner flipper field while the neutron is transversing
the coil (Larmor precession). The magnitude of this field may be tuned to turn
the neutron spin, for example, by IT so that the neutron polarization is inverted
with respect to the guide field direction when the neutron exits the flipper. An
extensive review about different types of flippers, for example radiofrequency
flippers, cryo flippers etc., may be found in (Williams, 1988).

3. 1. 3 The Polarization of the Scattered Beam


For the characterization of magnetic materials the analysis of the polarization
of the scattered beam is an essential part of the scattering experiment. If the
beam incident on the sample is polarized, performing this analysis allows one
to separate nuclear and magnetic scattering. We now elaborate on this point a
bit further.
The nuclear scattering potential II nuci sums the individual b 's, which for
nuclei with spin contain two parts:
" 1 "
b = A + "2Bs 1
1

where A and B are isotope-specific numbers, sand i are the spin operators of
neutron and nucleus, respectively. The spin-independent part of the potential
gives rise to scattering without spin flip, because the neutron spin is not
involved (usually referred to as nuclear scattering). The spin-spin interaction
gives rise to scattering without spin flip for the component of i parallel to s
whereas the two components perpendicular to s will scatter with spin flip
(Moon et al., 1969). Usually the nuclear spins are not ordered and this
contribution to the scattering (hence called spin-incoherent scattering) is 2/3
with spin flip and 1/3 without spin flip. The same rules apply for scattering
from a magnetic moment Jl: scattering from the component parallel to s is
without spin flip whereas the two components perpendicular to s will scatter
with spin flip. However, as we wrote in Section 3. 1. 1, only components of 1..1
perpendicular to q are effective.
Given an initial beam polarization p, we may summarize the effect of the
different types of scattering on the scattered beam polarization pI as follows:
Nuclear scattering: pI = P.

Spin-incoherent scattering: pI = - i
P.
78 G. Ehlers, F. Klose

Magnetic scattering from macroscopically isotropic magnetic systems:

p' =- q(q P)
q2 .

The last formula was first obtained by Halpern and Johnson in their classic
paper on magnetic neutron scattering (Halpern and Johnson, 1939). Its
meaning is: The scattered beam is polarized in the direction of the scattering
vector, the spin is flipped, and the polarization also depends on the relative
orientation of q and P. However, the formula is not universal. It holds only for
macroscopically isotropic cases, for example paramagnets, antiferromagnets,
generally powder samples, and (non-magnetized) multidomain ferromagnets
(as long as they do not depolarize the beam), but not for macroscopic single
crystals with non-collinear magnetic order.
Figure 3.5 shows a classical example of how paramagnetic moments can
be measured (Moon et aI., 1969). The substance used was a MnF 2 powder.
The unpolarized beam data set (upper panel) shows Bragg peaks at various
scattering angles together with a distinct background signal that falls off
towards high q. In the polarized beam measurements (P was kept parallel to
q), the Bragg peaks appear only in the non-spin-flip channel (middle). This
clearly verifies their non-magnetic origin. The spin-flip channel contains solely
the paramagnetic scattering signal (lower panel), decaying corresponding to
the characteristic form factor of the Mn 2 + ion.
In the most general case, when the Halpern/Johnson formula can not be
used, we have to write
p' = sp + P"
where S is called the polarization tensor and P" is the polarization created by
the scattering process. S has (at most) six independent elements. These can
be measured if the magnetic sample is kept in zero external magnetic field.
The corresponding technique is called spherical polarization analysis. The
requirement of zero sample field can readily be seen: While the incident
polarization P would be parallel to the sample field, any component of the
scattered polarization p' perpendicular to the sample field would precess
around the field and the beam average of those components would be zero.
The proper spectrometer operation of course requires a magnetic guide field,
but it has been possible to construct a zero-field sample chamber. This device
is called CRYOPAD (Tasset et aI., 1999). An experiment consists of
measuring the scattered polarization with the incident polarization along each
of the axes x, y, and z. For each of the three incident polarizations, the
scattered polarization is analyzed along all three spatial directions. The
required formulae for the individual elements of Sand P" can be found in the
literature in a handy form (Brown, 2001). Since the experimental setup is
difficult, and the diagonal elements of S can be measured in a different way
(see below), the technique of spherical polarization analysis is only applied if
Characterization of Magnetic Materials by Means of Neutron Scattering 79

5000

MnF z
4000
c:
's
M
... 3000
'"0-
<Il
Unpolarized beam
c:
2000
~
Z'" ~
1000
~
....... ~ -.. Ut.l
0
2000

1600
Polarized beam with

1200 -- --
polarization analysis
PIIK

c: 800 I
's Flipper off
0
N
... 400
~
'0"-
<Il
c: .... A.

I\..~ '-J tl
~
Z'"
0
600
"'
....
..
400

200
c-y....... ~
Flipper on
....
~....,_
~

----
'9
---..
--=--slclZground- - - -
0 10 20 30
Scattering angle (' )

Figure 3. 5 Separation of paramagnetic scattering through polarization analysis, Upper


panel: unpolarized beam measurement; middle: polarized beam measurement (non-spin-
flip channel); lower panel: spin-flip channel. Note that the polarized beam data have much
lower absolute count rates compared to the unpolarized beam measurement because of
intensity losses associated with the polarizer/analyzer equipment (Moon et al.. 1969),

S has off-diagonal elements, Such elements may appear in magnetic crystals


with non-collinear magnetic order,
An experimental setup that uses a magnetic guide field at the sample
position is experimentally easier to realize and allows one to measure the
three diagonal terms of S, This technique is called three-directional
polarization analysis (Scharpf and Capellmann, 1993). The magnetic sample
field is applied along each of the axes x, y, and z, and one measures spin
flip and non-spin flip cross sections for all three directions (for more details,
see the discussion related to the D7 instrument in Section 3, 2, 1), Nuclear and
80 G. Ehlers, F. Klose

spin-incoherent neutron scattering always appear in the diagonal elements of


S. because P and p' are parallel. Magnetic scattering gives rise to diagonal
elements of S. as long as the Halpern Johnson formula can be applied (see
above). Since nuclear. spin-incoherent and magnetic scattering affect the
polarization in different ways. one can combine the six measured cross
sections to give any of the three contributions separately.

3.2 Elastic Magnetic Scattering

3. 2. 1 Small-Angle Scattering
Small-angle neutron scattering (SANS) is extensively used for investigating
nanostructured materials. It probes length scales in the nanometer regime and
thus can be used to study microstructural features of matter such as size.
shape and magnetization of precipitates. As in other diffraction experiments.
the SANS intensity is measured as a function of the momentum (more precisely
wave vector) that is transferred from neutron to sample during the scattering
process. i. e . as a function of the scattering vector q:
q = k; - k,.
The scattering vector q is simply the difference between the neutron's incident
wave vector k; and its scattered wave vector k,. In SANS experiments. the
scattering is assumed to be elastic. i. e.. I k; I = I k, I = 21T/;. and I q I =
41T sin(e)/;.. where;' is the neutron's wavelength and 28 is the scattering
angle. The momentum transfer is inversely proportional to the length scale of
investigation. The available q range of SANS instruments is between 10- 1 A-I
and 10- 4 A-I corresponding to mesoscopic object sizes in the range between
approximately 10 to 10000 A. In contrast. in the diffraction regime, q is on the
order of 1 A-I and therefore suitable for probing interatomic distances.
The intensity is obtained by summing up the scattering amplitudes of all
atoms in the sample. weighted by the phase shift at each atomic position r:

/(q) = Ifb(r)e;Qrd3rI2

where the integral extends over the entire sample and b (r) is the local
(coherent) scattering length density. For magnetic materials. the latter has
magnetic contributions. As discussed in Section 3. 1. not only the magnitude of
the local magnetic moment but also its orientation relative to the incoming
neutron polarization playa role in the scattering process.
Characterization of Magnetic Materials by Means of Neutron Scattering 81

Small-angle scattering signals resulting from the magnetic microstructure


can be separated from those resulting from variations in the atomic density or
secondary phases. For unpolarized neutron beam experiments, this Qan be
done by investigating the magnetic field dependence of the SANS signal. Using
polarized neutrons, one can extract magnetic information by variation of the
incident neutron beam polarization in a fixed external field (Wiedenmann,
2001).
The following is an excellent example of the usefulness of SANS
measurements using unpolarized neutron beams. Weissmuller et al. recently
investigated the magnetic properties of nanocrystalline Co and Ni samples and
demonstrated that SANS measurements are able to provide quantitative
information on CD the magnetic microstructure, ~ the exchange stiffness
constant and @ the magnitude and microstructure of the magnetic anisotropy of
these materials (Weissmuller et aI., 2001).
Nanocrystalline ferromagnets have interesting magnetic properties due to
their strongly reduced grain size d. As the latter reaches nanometer
dimensions, random jumps in the local orientation of the magnetic easy axis
are introduced, leading to altered magnetic properties compared to ordinary
polycrystalline bulk materials. The effect of the reduced grain size on the
magnetic properties critically depends on the magnitude of the grain size
relative to a magnetic exchange length I K = VA / K, where A is the
ferromagnetic exchange-stiffness constant and K an anisotropy energy
coefficient. Nanocrystalline hard magnets have I K < d , with the
magnetization axis locked onto the easy axis of each grain and an enhanced
remanence due to gradients in the orientation of the magnetization at grain
boundaries. By contrast, nanocrystalline soft magnets have I K > d. In this
case the magnetization cannot follow the changes in the orientation of the easy
axes on the scale of the grain size. Instead, the local magnetization direction
results from an effective averaging of the anisotropy in many neighboring
grains. As the work of Weissmueller et al. (2001) demonstrates, SANS is an
excellent tool for measuring magnetic structures in these length scales.
The nanocrystalline Co samples that are the subject of the following
discussion were prepared by pulsed electrodeposition. Although the mass
density was practically identical with that of bulk Co, the structural grain size
was found to be quite small, about 10 nm, and the mean distance between
stacking faults was even smaller, only about 2 nm. Since the volume fraction
of structural inhomogeneities is small, the ratio of magnetic scattering signals
to the magnetism-independent nuclear scattering backgrounds in the SANS
data were relatively high, allowing this effect to be observed.
Figure 3.6 compares magnetization and magnetic SANS data of the
nanocrystalline Co sample. The hysteresis loop measurements (Fig. 3.6a)
indicate that the material apparently saturates in fields of about O. 5 T. The
magnetization data are practically identical for low-temperature (10K) and
ambient temperature measurements. Figure 3. 6b shows that SANS
82 G. Ehlers, F. Klose

1500 10 5 . ncCo
T=295 K
1000 10 4

500 103

]' T.
0
</>
T 102
~ E
~
~ -500 q 10 1

nc Co
"hi
-1000 " 10
-0- T=10 K
I - T=300K 10- 1
-1500 i.I=::::::=~~~L_~=~
-2 -I 0 2
/loH (T) 0.02 0.1
q (nm-I )
(a) Magnetization (b) SANS

Figure 3.6 (a) Magnetization isotherms at T= 10 and 300 K for nanocrystalline Co. (b)
Experimental differential scattering cross section doL / dO versus modulus q of the
scattering vector for magnetic fields of (from top to bottom) 5, 43, 87, 180,390,770,
1140 and 1740 mT (Weissmuller et aI., 2001).

measurements, which in this case essentially probe the spatial homogeneity of


the magnetic moment orientations, reveal complementary information: In
contrast to the seemingly saturated magnetization, there are still substantial
changes in the SANS signal as the magnetic field is increased beyond O. 5 T.
Due to this strong field-dependence it can be excluded that the scattering
originates from magnetized particles in a nonmagnetic matrix or from
nonmagnetic particles or pores in a magnetic matrix. The scattering contrast
between a particle and the matrix would remain essentially constant,
independent of H, once the sample was near saturation. Therefore, the
authors argue that the scattering must result from a continuous, small and
periodic variation of the spin misalignment angle relative to the overall
direction of the field. This hypothesis is supported by the fact that the
curvature in the log-log plots of differential scattering cross section d~ / dO
versus q is shifted to larger q as the field strength is increased. In SANS data
the scattering vector at maximum curvature often corresponds to 2rr over a
characteristic length scale; therefore the observation is consistent with the
notion that increasing the magnetic field leads to a suppression of the long-
wavelength magnetic fluctuations of the magnetization, so that the dominant
wavelength is progressively reduced.
Another important result of these SANS measurements is that other factors
besides the grain size d can significantly affect the magnetic microstructure.
Examples of such factors are twin boundaries and centers of strong anisotropy
or of antiferromagnetic coupling, potentially due to changes in the atomic
coordination and interatomic spacing in the core of grain boundaries or
Characterization of Magnetic Materials by Means of Neutron Scattering 83

dislocations.
In general, magnetization distributions can be studied with unpolarized
neutron beams (i. e., with randomly oriented spins) if the scattering intensities
from chemical and magnetic structures are in the same order of magnitude.
The utilization of polarized neutrons has one major advantage: it makes it
possible to modify the relative contrast between magnetic and non-magnetic
particles, amplifying weak magnetic signals that may be shadowed by strong
scattering from other sources. Polarized SANS does not necessarily need to
include polarization analysis of the scattered neutrons, although this can be
done for example by using a 3 He filter system (see Section 3. 1.2). In certain
situations, polarization analysis reveals further details regarding the relative
orientations of the magnetic domains in the sample.
Because of experimental difficulties and a strong reduction in neutron flux
that is inherent to polarized beam techniques, only a few of the ~ 25 SANS
instruments in the world have a polarized beam option available. As an
example, Fig. 3.7 shows the schematic layout of the polarized neutron SANS
instrument at Hahn Meitner Institute Berlin (Keller et al., 2000).
Velocity Transmission Magnetic C n' S . fr Area
selector polariser guidefield 0 unators pm lpper detector

l-16m
16 m,12 m,8 m, 2 m,l m

Figure 3.7 Polarized SANS instrument V4 at Hahn Meitner Institute Berlin (figure taken
from HMI web site) .

Neutrons enter from the left and are monochromatized (A i\ / i\ ~ 10% )


using a velocity selector. Beam polarization is achieved by using a wedge-
shaped supermirror transmission polarizer which can be rotated into the beam.
This kind of polarizer has a distinct advantage, particularly for SANS
applications: since it uses only the transmitted neutrons, it does not deflect the
beam and therefore does not require a realignment of the 30 m long
instrument. The polarizer is installed in the first section of the 16 m long
collimator drum. The collimation length is variable (minimum: 1 m, maximum:
12 m) in order to achieve variable angular resolution. Collimation sections that
are not in use are replaced by Ni neutron guide sections via a drum
84 G. Ehlers, F. Klose

mechanism. The beam size and collimator aperture openings are 3 cm x 5 cm.
A magnetic guide field of about lOG is needed between the polarizer and the
sample to avoid depolarization of the neutron beam. A radio-frequency spin
flipper. consisting of a longitudinal ac magnet field coil and a static gradient
magnet field. is installed in front of the sample to allow measurements with
both neutron polarizations. A particular feature of this spin flipper design is
that no material is needed in the path of the beam. This is important for high-
resolution SANS applications because small-angle scattering from the wire
materials that are used in the usual Mezei flipper design (Fig 3. 7) would
seriously degrade the angular resolution of the instrument. In order to cover
different q ranges, the two-dimensional 3 He-detector with 64 x 64 elements of
1 cm xl cm can be positioned at any distance between 1 and 16 m from the
sample in the horizontal direction. In the polarized neutron mode, the V4
instrument achieves> 30 % of the corresponding unpolarized neutron flux at an
average beam polarization of """gO % .

3.2.2 Neutron Diffraction

Neutron diffraction is an extremely powerful technique for the determination of


magnetic ordering in materials on an atomic scale. Of all types of neutron
scattering instruments in the world, diffractometers are by far the most
common. The ordering pattern of the spins and the size of the ordered
magnetic moment can be obtained from the data in a straightforward way
unless the ordering is extremely complicated and exotic (only neutron
diffraction has taught us that those cases exist). Many diffractometers have
been built which are optimized for magnetic scattering. Surprisingly, not many
of them actually use polarized neutrons. The reasons are that CD polarizing the
beam significantly reduces the beam intensity, and ~ in most cases il is
sufficient to observe the additional scattering in reciprocal space while cooling
the sample below the ordering temperature. In particular this is true for
powder samples, where one averages any way over all equivalent crystal
directions. A microscopic model of the magnetic ordering can be developed
that is compatible with the data, and in many cases the choice of the model
will be unique. The symmetry of the spin structure is usually revealed by the
presence or absence of specific reflections. For example. a collinear structure
with moments al igned parallel to a (001) crystal axis has the signature of
extinct (001) reflections, because only spin components perpendicular to q
are relevant for magnetic scattering.
There are, of course, cases where a polarized beam is indispensable for
the determination of the magnetic structure of a material. These are usually
single crystals with either a complicated ordering pattern (non-collinear
ordering, chemically different magnetic ions in one unit cell, etc.) or with a
highly symmetric ordering which has only few reflections. The latter case may
Characterization of Magnetic Materials by Means of Neutron Scattering 85

cause a problem because towards high q, intensity is increasingly suppressed


by the magnetic form factor.
One can distinguish two types of diffractometers which are dedicated to
using polarized neutrons. A diffractometer such as 03 at ILL, Grenoble,
France, shown in Fig. 3.8 (see also http://www.ill.fr/YeliowBook/03) is
built to study single crystals. In the past it has been used for the quantitative
determination of magnetization distributions of magnetically ordered single
crystals (Ressouche et aI., 1993), for the investigation of exotic
antiferromagnetic structures (Hiess et al., 2001), for the determination of
antiferromagnetic form factors (Brown et al., 1999), and the search for spin
liquids or short-range ordering in paramagnetic phases. 03 can host the
CRYOPAO device (Tasset et aI., 1999) which allows one to do spherical
polarization analysis. 03 mostly uses hot neutrons and has COO 92 Feoo8 and
Heusler alloy CU2 MnAI monochromator crystals. A 3 He neutron spin filter is
currently being installed to analyze the scattered polarization which is
significantly better than using a spin analyzer crystal.
Helium gaz
pumping group
I

Liquid nitrogen
dewar

Cryomagnet

Polarising
monochronator
Harmonic filters,
secondary shutters
and monitor
Motorised slits Rotating platform on air-pads

Figure 3.8 Schematic layout of the D3 diffractometer at the ILL (from D3 web site) .

A complementary type of diffractometer is 07 at ILL, Grenoble, France


(see www.ill.fr/YeliowBook/07).This instrument is optimized for studying
diffuse magnetic scattering (Stewart et al., 2000; Cywinski et al., 1999).
Figures 3. 9 and 3. 10 show the general layout. A doubly-focusing graphite
monochromator is used to define the incident energy. The beam then passes
86 G. Ehlers, F. Klose

through a supermirror polarizer, a Mezei spin fl ipper and a disk chopper (the
chopper is optional and allows for performing energy analysis of the scattered
neutrons using the time-of-flight method). A removable set of six coils, wide
enough for a furnace or orange cryostat to enter, is placed around the sample
position in order to define the incident beam polarization direction (Fig. 3.9).
D7 has four detector banks placed at 1. 5 m distance from the sample. Each of
the 32 single 3 He detectors has an individual supermirror spin analyzer. The
necessary guide field between sample and analyzers is provided by permanent
magnets. The instrument is flexible: Guide field and spin analyzers can be
replaced by a radial collimator for measurements without polarization analysis.
The following modes of measurement are possible:
(1) No polarization analysis: unpolarized neutrons in-no analysis after
scattering;
(2) Polarized Neutron Diffraction: polarized neutrons in-no analysis after
scattering;

Beryllium filter

l----'Od--\+-+ AnaIyzers
(removed)

Figure 3. 9 General layout of the D7 polarized neutron diffractometer. The monochromatized


neutron beam enters the instrument from the upper right corner through a beryllium filter that
removes the unwanted higher order reflections of the monochromator. The sample is
located in the center position inside a cryostat and is surrounded by four detector banks.
The latter are mounted on air cushions to allow for variable scattering angle coverage.
Characterization of Magnetic Materials by Means of Neutron Scattering 87

(3) Z polarization analysis: two measurements (with and without spin-


flip); allows the separation of coherent and incoherent scattering in non-
magnetic systems;
(4) XYZ polarization analysis: six measurements; allows the separation
into coherent, incoherent and magnetic contributions.
D7 is not only versatile for magnetic studies but is also used by other
communities, because polarization analysis allows for separating coherent
from spin incoherent scattering, as for example, in polymers (Eilhard et al.,
1999) or liquids (Garcia-Hernandez et al., 1999). A general overview of the
applications of elastic scattering of polarized neutrons to study nonmagnetic
materials is given by Gabrys (Gabrys, 1999).
Unpolarised "white"
beam
Monochromator

Unpolarised
monochromatic
neutron beam
1/,.-\/ \

Supermirror
polariser
Neutrons polarised
Supermirror by supermirrors
analyser FliPper~~I,--t'--.:.\'--.:.\r=\~t\~\~\~\d..------,
_ _-, Neutrons flipped
by flipper
HHH,H
Chopper

Sample

Some neutrons
flipped by sample
HHHHI
Neutrons analysed
Detectors by supermirrors
t\ t t

Figure 3.10 Schematic layout (top view) of D7 for polarized neutron scattering
experiments with polarization analysis (see text). The diagram shows only one of the four
spin analyzing detector banks.

As an example for magnetic diffraction, we review the study on atomic


and magnetic correlations in an archetypal spin glass CuMn, performed by
neutron scattering on single crystals.
88 G. Ehlers, F. Klose

The term "spin glass" refers to the particular magnetic state of a system
of spins with two basic ingredients: CD spatial disorder and (Z) frustration due
to competing antiferromagnetic (AFM) and ferromagnetic (FM) interactions.
Spatial disorder is, for example, realized in dilute solid solutions of magnetic
ions such as Mn or Fe in noble metals (CuMn, AgMn, AuFe are well studied
examples), or by partially substituting a magnetic ion in a stoichiometric
crystal (as in Eu x Sr 1- x S). The magnetic interactions can be of different
origin. In non-metallic rare-earth compounds, the 4f electrons couple via direct
isotropic Heisenberg exchange (EuS). In the case of 4f ions in metals the
interaction is thought to be based on the Ruderman-Kittel-Kasuya-Yosida
(RKKY) mechanism (Ruderman and Kittel, 1954; Kasuya, 1956; Yosida,
1957). The 6s conduction electrons (and in certain cases electrons in the
partially overlapping 5d shells) are polarized by localized 4f moments and
neighboring 4f moments, in turn, are coupled via the polarized conduction
band (it has to be said though, that the original RKKY model is far too simple
and has seen numerous attempts to extend it to real istic cases). The main
feature of the RKKY interaction is that it causes an oscillatory dependence of
the coupling constant between two sites as a function of distance. In metallic
spin glasses with 3d moments, the situation is very complicated and not fully
understood. At low concentration of magnetic impurity (roughly below 10%),
the moments are, to a good extent, localized and an RKKY type interaction
can be assumed, but at high concentration the 3d moments become itinerant.
It depends on how the magnetic ions are distributed and how the 3d states
overlap. Magnetic anisotropy may also play a decisive role for the
understanding of some systems.
Experimentally, the following phenomena are characteristic for spin
glasses: CD a peak in the ac susceptibility Xac (T) at low temperature, the
position of the maximum defining the spin glass temperature T f , (Z) absence of
magnetic Bragg peaks in neutron diffraction, meaning that the spins "freeze"
randomly without long range order, @ there is no anomaly in the magnetic
specific heat at T f but a broad bump at higher temperature (at about 1.3 T f ) ,
the magnetization shows a strong history dependence below T, and a slow
non-exponential relaxation after perturbation.
In the CuMn system the situation gets even more complicated because
atomic short range order is observed: a Mn atom is preferentially surrounded
by Cu atoms. This is important as it has an influence on the distribution of Mn-
Mn distances in a sample. Again, polarized neutrons give unique insight into
this particular problem on an atomic scale, because one can look at magnetic
and nuclear correlations at the same time in the same sample.
Neutron diffraction has contributed to this particular field from the very
beginning (Meneghetti and Sidhu, 1957>, showing the absence of long range
magnetic order, unless for a high Mn concentration above 70 at. %. These
early measurements did reveal broad magnetic "humps" situated between the
( 100) and (110) Bragg peaks, indicating short range ferromagnetic order.
Characterization of Magnetic Materials by Means of Neutron Scattering 89

The centers of these humps were later identified as the (1, 1/2, 0) position
(and equivalent positions) in reciprocal space. Experiments on single crystals
later showed the existence of additional and somewhat sharper reflections at
(1, 1/2 <5, 0) (Cable et al., 1982; Lamelas et ai., 1995), which were
attributed to" spin density wave" (SOW) like AFM correlations. The
parameter <5 depends on the Mn concentration and could be followed to very
dilute systems with 1 at. % Mn and even less. An extrapolation to zero Mn
content gave a value for the position of these peaks well correlated with the
Fermi wave vector 2 k F of Cu. The proposed interpretation of these findings
was that CuMn alloys are incommensurate SOW antiferromagnets, with the
magnetic order driven by the intrinsic instability on the Fermi surface of Cu
(Fig. 3. 11). This conclusion opposes the widely accepted view of CuMn alloys
as spin glasses.
Motivated by this controversy, a recent study on a single crystal of Cu-
5at. % Mn was performed at 07 at ILL (Murani et ai., 1999; Stewart et ai.,
2000) .

CUO.95 M110.05

Q!.

(b)

Figure 3. 11 Nuclear (a) and magnetic correlations (b) simultaneously measured by


polarized neutron diffraction in a Cu-5 at. % Mn single crystal at low temperature (, . 5 K) .
Data taken from (Stewart et al., 2000).

These are probably the most precise elastic data available on CuMn. The
earl ier experimental findings are all confirmed: CD atomic short range order is
seen by the broad intensity maxima around (1, 1/2, 0) and (1/2, 1, 0) in the
nuclear correlation map, (?) both the broad magnetic correlation peak at the
same positions as well as the SOW peaks at incommensurate positions are
found. Nevertheless the above interpretation is not forced, because these
magnetic features actually represent only a small fraction of the total magnetic
scattering intensity. One also observes a quasiuniform background present at
all q, originating from the random freezing of most of the Mn spins. In
addition, as comparison to other available data shows that this "background"
90 G. Ehlers, F. Klose

becomes more important as the Mn concentration decreases. Another strong


argument supporting the spin glass view comes from the observation (Mezei
et al., 2000) that the intensity of this scattering is very much temperature
independent up to 100 K (the quasielastic width of the scattering function
increases of course a lot with temperature) and its q = 0 limit corresponds
well to the bulk susceptibility measured above T j Therefore it can be
concluded that SDW like AFM correlations coexist with the spin glass state but
are not dominating the magnetic behavior of CuMn crystals at low
temperature.

3. 2. 3 Reflection of Neutrons from Magnetic Surfaces


and Interfaces
Polarized neutron reflectometry (PNR) has a successful track record of
providing unique insights in problems of magnetic surfaces, thin films,
interfaces, and multilayer systems (for a recent review, see for example
(Ankner and, Felcher, 1999. Areas of current fundamental science
addressed by this technique include flux penetration and flux-lattice ordering in
superconductors, nucleation and growth of structured surfaces, magnetic
moment formation in thin films, interface polarization, interfacial coupling and
quantum confinement, giant and colossal magnetoresistance. In the future,
investigations on magnetic domains and patterned structures of magnetic dots
or other nanoparticles, self-assembled layers and integrated materials such as
polymers combined with magnetic materials, molecular magnets, etc. will
become increasingly important. Fundamentally new scientific insights gained
by PNR were and will be important in the development of future thin-film based
applications, such as new hard and soft magnetic materials to improve the
efficiency of energy delivery systems (e. g., motors, transformers, etc.)
(Fullerton et al., 1999), magnetic recording media and magnetic sensors for
computers (Speriosu et aI., 2000; Thompson and Best, 2000), new magnetic
memory technologies such as non-volatile magnetic random access memory
(MRAM) and other so-called spintronics devices (Prinz, 1999).
Next generation instruments with much higher available neutron flux such
as the Magnetism Reflectometer at the Spallation Neutron Source (Lee and
Klose, 2001) will provide unprecedented experimental capabilities. Most
importantly, they will be capable of routinely detecting weak off-specular
scattering signals resulting from chemical/magnetic structures within the layer
plane using off-specular/ grazing-incidence small-angle scattering (GISANS)
techniques (Mueller-Buschbaum et aI., 1999; Felcher and te Velthuis, 2001).
Such experiments are unreasonably slow on today's instruments. The SNS
reflectometers will be the first neutron scattering instruments capable of
directly detecting scattering signals (finite size oscillations, see Fig. 3. 13)
from monolayer films. The high neutron flux will make possible in-situ structural
Characterization of Magnetic Materials by Means of Neutron Scattering 91

or magnetic phase-diagram determinations as functions of thermodynamic


parameters such as temperature, pressure, atmosphere, magnetic field, etc.
and will even provide the ability for time-dependent studies (e. g., pulsed
magnetic, electric, light or other fields, etc.) The availability of polarized
neutrons and the polarization analysis capability suggest that the instrument
will also be used for specific studies on non-magnetic thin film samples.
Examples of the latter cases include contrast variation, incoherent background
reduction and phase determination for direct inversion of reflectivity data into
real-space scattering-length density profi les (Schreyer et al., 1999).
Figure 3. 12 schematically displays the scattering geometry in PNR
experiments. Note that a similar geometry also applies for general diffraction
experiments with polarization analysis. A peculiarity of the thin film case is
that the shape anisotropy usually forces the magnetic moments to lie within the
plane of the surface. The lateral arrangement of the moments, however, is
generally determined by other causes, for example resulting from crystalline
anisotropies or exchange coupling energies. Therefore, in PNR experiments,
magnetic fields B ext are typically applied within the sample plane (e. g., along
the y axis in Fig. 3. 12). The direction of B ext also defines the polarization axis
of the neutron beam. In polarized reflectometry experiments, one measures
four different cross-sections as functions of the scattering vector q. The
scattering vector can be scanned in two different ways: reactor type
instruments typically use monochromatized neutrons and vary the scattering
angle, while spallation neutron source instruments use the time-of-flight method
at a fixed angle of incidence to scan A and consequently q. R + + and R - - are
the non-spin flip reflectivities (the first superscript characterizes the incident
neutron polarization and the second the exit polarization; + corresponds to
"spin up", - to "spin down", respectively). These cross-sections are
sensitive to the chemical layering of the film structure as well as the magnetic
moment component Mil that is oriented parallel to the neutron polarization

yt
r

Figure 3. 12 Schematic representation of the scattering geometry in polarized neutron


reflectometry experiments (see text). The thin film sample is usually deposited on a
substrate material. The magnetic moment vector M lal may have components parallel and
perpendicular to the neutron polarization axis.
92 G. Ehlers, F. Klose

axis. Provided that no other causes are present that lead to spin-flip scattering
(see Section 3. 1.3), R - + and R + - are sensitive to magnetic moment
components M~ that are perpendicular to the neutron polarization axis (Zabel,
1994a). Therefore, PNR reveals not only the depth profile of collinear
magnetic structures, but also allows studies of non-collinear magnetic
arrangements, including chiral structures (O'Donovan et al., 2002).
Provided that the film is laterally homogeneous, the scattering is
specular, i. e., the angle of incidence ai equals the take-off angle at. In this
case, reflectometry measures the chemical and magnetic depth profile along
the z axis. In the first order Born approximation, the scattered intensity is
given by (Als Nielsen, 1986)

J(qz) cc ~~ If d~~Z)exp(_ iqzz)dz 1


2

where V (z) is the potential that the neutron experiences at a depth z inside
the film. This means that the specular intensity is proportional to the Fourier
transformation of the gradient of the potential profile perpendicular to the
surface. For a homogeneous layer consisting of a pure element, and with all its
magnetic moments aligned parallel to B ext ' V (z) is constant throughout its
depth and is given by

V = 21T hn (b p) .
mn
In this expression, m n is the neutron mass, n is the number density and b + P
is the nuclear, magnetic scattering length of the atom species, respectively.
For spin up neutrons, the magnetic contribution p has to be added to b, while
for spin down neutrons it has to be subtracted. For this particular case of
aligned magnetic moments only R++ and R-- will have intensity. Since no
perpendicular components of magnetic moments are present, no spin-flip
scattering wi II occur and R - + = R + - = O. Note that, in contrast to the
diffraction regime, V (z) is a locally averaged potential. In the case that the
material consists of different atomic species, their individual contributions to
the average local potential have to be summed up. For a sequence of layers,
V(z) changes continuously along the depth of the film.
It is well known that the Born approximation fails for qz- 0 (the intensity
would become infinite at qz = 0). In this regime, optical methods need to be
applied that correspond to full dynamical theory taking multiple scattering,
refraction and absorption effects into account (Majkrzak, 1991; Zabel,
1994b). A particularly useful method is the recursion scheme of Parratt
(Parratt, 1954). This method does not contain approximations and provides
exact solutions for reflectivity profiles. Interface roughness can be simulated
by slicing the interface region in arbitrarily small regions that approximate the
gradient of the potential.
Figure 3. 13 displays PNR spectra of typical thin film structures. The
Characterization of Magnetic Materials by Means of Neutron Scattering 93

reflectivity functions have been calculated using the Parratt 32 program


developed by C. Braun, Hahn-Meitner Institute Berlin. For all film structures,
the region of total reflection and the overall intensity is different for the R + +
and R - cross section. While the reflectivity functions of the Fe surface are
featureless declining proportional to 1/ q4, the interference between partial
waves reflected by the vacuum/film and film/substrate interfaces causes an
oscillatory behavior of the reflectivity for the 500 A Fe film (Kiessig fringes).
The main feature of the multilayer reflectivity function is that strong Bragg
peaks occur, which result from the 80 A double-layer thickness (50 A Fe +
30 A Nb).
10 , - - - - - - - - - - - - - - - - ,
10-1 Fe surface
10-2
10-3
10-4
10--5
10-6 ' - - - - ' - - - - - ' - - - - - ' - - - - - - - " : = . . - - - '
C 10==------------,
:~ 10-1 500AFeon ~ ~
~
.;: 10-2 Si substrate 500 A Fe
~ 10-3
10-4 Si
~ 10-5
OJ
;z: 10-6 '--_-'--_--'-_---'--_-"-""-'-":>...AJ
10 ,---,.--------,,.------;:------,
10-1 ~ ~30~Nb
10-2 50AFe
(X 12)
10-3
10-4
10-5 Si
10-6
0 0.04 0.08 0.12 0.16 0.20
q (A.-I)

Figure 3. 13 Calculated PNR spectra for different layer sequences. Upper data set: a
bulk Fe surface; middle: a 500 A Fe film on Si substrate; bottom: a multilayer consisting
of 30 Nb/50 A Fe with 12 repetitions on Si substrate. In all cases it is assumed that the
magnetic layers are saturated along the magnetic guide field B exl' R + + is plotted in blue
and R - - in red color.

As mentioned above, laterally structured magnetic films become an


increasingly more important research topic. Figure 3. 14 summarizes a neutron
scattering experiment on a patterned magnetic film (Lee et al., 2002a). The
data were collected using the POSY2 time-of-flight neutron reflectometer at
Argonne National Laboratory. In this experiment, the sample consisted of
lithographically produced rectangular Ni stripes of 10 IJm x 2 IJm area and 100A
thickness that are separated by 2IJm from the neighboring stripes. A Si wafer
serves as substrate. The scattering pattern displayed on the right side is
typical for this kind of sample and demonstrates the usefulness of the
94 G. Ehlers, F. Klose

technique. It shows specularly reflected intensity, where the final scattering


angle Bj is constant with wavelength and additional off-specular intensity that is
located below and above the specular intensity. Note that, since the substrate
is largely transparent for neutrons, also a transmitted refracted beam is visible
below the scattering horizon. Analyzing such scattering patterns reveals
information on size, shape and periodicity of these nanostructures as well as
depth-resolution (Toperverg et al., 2000; Toperverg, 2001). Of course, this
is particularly useful if the surface pattern is not known from the outset, for
example in the case of self-assembled structures. Magnetic information also
can be obtained by using polarized neutrons and polarization analysis. For
example, for the case of the Ni stripes, it allows to determine details of the
magnetization reversal process. In particular, it permits distinguishing
between domain reversal by a rotation or along a particular axis along the
surface (Temst et al., 2001).

I Incident I

--------~~~~~

Sample:
IOmx2m Wavelength 2
Ni stripes
(l00 A film
thickness)

Figure 3. 14 Neutron reflectivity in time-ol-flight mode on a patterned magnetic film


structure. The left side shows the experimental arrangement and schematically the
patterned surface in side and top view. On the right, the corresponding scattering pattern
(contour intensity plot of final scattering angle vs. neutron wavelength) is displayed (Lee
et al., 2002a).

If the lateral length scale of the surface structure is large compared to the
corresponding projection of the coherence length of the neutron beam ( the
latter is typically on the order of microns to tens of microns), there is
practically no off-specular intensity and no information on the size of lateral
structure is available. Nevertheless, polarized neutron specular reflections may
contain information on the lateral arrangements of large-scale magnetization
inhomogeneities, such as large domains. Lee et al. recently demonstrated this
approach and measured the field dependence of the ferromagnetic domain
dispersion X20f an exchange biased CoO/Co bilayer film (Lee et aI., 2002b).
Characterization of Magnetic Materials by Means of Neutron Scattering 95

In the following, we describe an experiment that clearly demonstrates the


advantages of using neutron scattering for characterizing magnetic phenomena
in surfaces and thin films. In this example, exchange coupling in magnetic
layered structures was studied. Exchange coupling, initially observed by
Grliberg et al. (1986) in Fe/Cr multilayers, is an interesting effect because of
its large impact on information technology due to the related giant
magnetoresistivity (GMR) effect, which has been found in several material
combinations. GMR is an extraordinarily large change (up to 100 %) of the
electrical resistance of antiferromagnetically coupled thin film structures upon
application of an external magnetic field (Baibich et ai., 1988). GMR
materials have quickly found various technical applications and are currently
used, for example, as advanced sensor elements in read heads of computer
hard disks.
Generally, exchange coupl ing of two magnetic layers through a non-
magnetic spacer layer can be described by using RKKY-Iike or quantum
interference models. Theoretically as well as experimentally, it is found that
the coupling energy J, which is a measure of the coupling strength, oscillates
approximately as

J cc (2k:ts)2sin(2kFts)

where k F and t s are the (effective) Fermi wave-vector and the thickness of
the non-magnetic layer, respectively. According to the formula, exchange
coupling is oscillating in sign, i. e., with increasing spacer layer thickness t s '
alternating ferro (FM)- and antiferromagnetic (AFM) coupling should be
observed. This dependence of the exchange energy on the spacer layer
thickness has been extensively studied in many systems, but surprisingly, only
very few papers exist in the literature dealing with a manipulation of the Fermi
wave-vector.
The main idea of the described experiment (graphically summarized in
Fig. 3. 15) was to use hydrogen absorption for changing the electronic
properties of the spacer layer material, including its Fermi wave-vector k F,
and demonstrate that this can result in a complete reversal of the magnetic
coupl ing (K lose et ai., 1997). Fe/Nb multilayers were chosen for these
experiments since oscillating FM/ AFM coupling had been positively identified
(Rehm et ai., 1997) and Nb, as the spacer layer, has a high solubility for
hydrogen (experiments with similar results have been carried out by
Hjorvarsson et al. (1997> on single-crystalline Fe/V superlattices) .
Figure 3. 16a shows neutron reflectivity data of Fe/Nb multi layers , each
with constant Fe thickness of 26 A but with four different Nb thicknesses.
These different chemical periodicities result in structural Bragg peaks as
indicated in the figure. The samples were measured in their virgin state in
"zero" external magnetic field because, due to the relatively small coupling
strength in the Fe/Nb system, even small magnetic fields would have affected the
96 G. Ehlers, F. Klose

measurements. Therefore, unpolarized neutrons were chosen which made the


use of a guide field unnecessary.

Figure 3. 15 Hydrogen charging of Fe/Nb multi layers and resulting reversal of the
magnetic coupling J

Structural

'.
ArM peaks peaks Coupling between :i Hydrogen
neigbboring
Fe layers
~
cL
co
v
to.

/
charging

---- Fe .g
v


---- Fe :::
bIJ
co
E
---- Fe
E
,~
cco

/
Hydrogen
---- Fe ~
co
---- Fe p
":I)
:::
outgasing
41.

v
---- re C
o 0.05 0.10 0.15 0.20 10 5 I 00 10 1 I 02 10 3
q (A-I) PII,(mbar)
(a) [26 AFelt ANb]Xn (b) [26 AFellS ANb]x 18
Figure 3.16 (a) Neutron reflectivity on Fe/Nb multi layers (26 A Fe/ t A Nb) x n (t is the
Nb layer thickness and n the repetition number). Note the extra half-order peaks at """ 15 A
Nb and""" 24 A Nb, which evidence anti parallel coupling of neighboring Fe layers (Rehm
et al., 1997). (b) Intensity of the antiferromagnetic Bragg peak of the multilayer (26 A Fe/
15 A Nb) x 18 (second data set from the top in (a)) as a function of the external hydrogen
pressure( Klose et aI., 1997).

Figure 3. 16a gives clear evidence that the magnetic coupl ing energy
oscillates in sign. With increasing Nb layer thickness, the extra half-order
peak resulting from coherent AFM coupling of neighboring Fe layers is visible
only for ~15 A Nb and ~24 A Nb, but not for ~8 A Nb and ~18 A Nb (in the
Characterization of Magnetic Materials by Means of Neutron Scattering 97

latter cases, intensity from ferromagnetic coupled Fe layers adds to the


intensities of the structural Bragg peaks) .
The multi layers were charged with hydrogen from the gas phase in a
vacuum chamber around the sample position of the neutron reflectometer at a
temperature of 473 K (the hydrogen concentration in the Nb layers is a function
the external hydrogen pressure). This allowed an in-situ observation of
changes of the magnetic coupling during the hydrogenation. Figure 3. 16b
shows the development of the AFM intensity for the (26 A Fe/ 15 A Nb) x 18
multilayer upon hydrogen absorption. It can be seen that the AFM intensity
(and therefore the AFM coupling of the Fe layers) is strongly suppressed with
increasing hydrogen content in the Nb layers indicating that FM coupling starts
to dominate. It is interesting to note that this effect is completely reversible
upon lowering the hydrogen concentration in the Nb layers by evacuation of the
hydrogen in the loading chamber. The observed effect can possibly be used to
fine-tune the magnetoresistive properties of GMR sensors or to build hydrogen
sensors based on the GMR effect.

3.3 Inelastic Magnetic Scattering

3.3.1 Studies of Elementary Excitations by Triple-axis Spectroscopy


Triple-axis spectroscopy (TAS) and chopper spectroscopy are both very
effective techniques for detecting spin dynamics in magnetic crystals (Shirane
et al., 2002). Important scientific areas that can be investigated using these
methods include strongly correlated electrons systems, high-temperature
superconductors, colossal magnetoresistive materials, quantum and molecular
magnetism, and itinerant magnets. Chopper spectrometers have been built at
reactor sources but are most effective at pulsed sources (Windsor, 1981, and
references therein). In the latter case, typically a fast spinning Fermi chopper
creates bursts of monochromatic neutrons. The energy of these neutrons can
be varied by changing the phase between the Fermi chopper and the
accelerator. Neutron energy changes during the scattering process are
detected by measuring the time-of-flight between sample and detector.
Chopper spectrometers can very effectively scan large areas in (q, w) space
using detectors with large sol id angle coverage. Polarization analysis,
however, is very difficult for this type of instrument. Only the recent progress
in the development of 3 He spin-analyzing cells (see Section 3. 1. 2) opens
realistic possibilities here.
The triple axis spectroscopy method can be favorably combined with
polarization techniques. An example for such instrument type is the IN20
98 G. Ehlers, F. Klose

spectrometer at ILL (Fig. 3 17).

Monochromator changer
Secondary shutter
Pyrolytic graphite filter

Figure 3. 17 Schematic layout of the IN20 triple axis spectrometer at ILL A combination
of Heusler-alloy (CU2 MnAI) (",) monochromator and analyzer (horizontally focusing) and
two Mezei-type spin flippers allows polarization and analysis of the incident and scattered
beams parallel or anti parallel to the vertical axis. The figure is taken from ILL web site
(http://www. ill fr/YeliowBook/IN20/).

In triple axis spectroscopy one compares the neutron momentum and


energy before and after the sample, determined on either side by Bragg
diffraction from monochromator and analyzer crystals. The principle is
schematically shown in Fig. 3. 18.

~
k"E,
n Beam
Detector

Figure 3. 18 General schematic layout of a triple axis spectrometer. M, S and A denote


monochromator, sample, and analyzer, respectively. The angle ljJ defines the crystal
orientation in the laboratory frame. E;, k; and E" k, are neutron energy and momentum
before and after the scattering process with the sample, respectively (for details see text).

The incident white beam is reflected (at an angle 2 em) from a focusing
monochromator crystal onto the sample. The neutron energy E, and momentum
k, incident on the sample depend via Bragg's law on lattice spacing used in the
monochromator and the angle em. Focusing is usually vertical to increase
Characterization of Magnetic Materials by Means of Neutron Scattering 99

intensity at the expense of vertical q-resolution. The choice of the


monochromator mainly depends on the wavelength, as was discussed in
Section 3. 1. 2. In the thermal and cold ranges, the (002) reflection from
pyrolytic graphite yields the highest intensity. Cold neutrons are usually
polarized in a second step by means of a supermirror polarizer. In the thermal
and hot ranges one also uses Si (111), and various reflections from Cu, such
as (111), (200) and (311). Heusler (111) crystals are used for direct beam
polarization (see Section 3. 1. 2), mainly in the thermal neutron range, as for
example at the IN 20 spectrometer (F ig. 3. 17). The beam scattered from the
sample at an angle 2 es is reflected from an analyzer crystal into the detector.
e
The angle a at the analyzer determines the neutron momentum k, and energy
E, after the sample. Collimators may be placed in the flight path to reduce the
angular divergence of the beam and improve q-resolution. Filters are often
used to reduce intensity of contaminant short wavelength neutrons.
Each experiment measures the scattering function S (q, w) along certain
lines in (q, w) space, where q = k; - k f and w = E; - E f In the horizontal
plane the scattering function has three independent parameters (wand two
horizontal components of q), while there are four adjustable observables: the
energies E; and E f , the scattering angle 2 es and the angle lIJ which defines the
crystal orientation in the horizontal plane of the laboratory frame. Usually one
keeps either E; or E f constant during a particular scan.
The accessible ranges of energy an? momentum transfer depend on the
incident wavelength: using cold neutrons, one can typically measure up to
hw = 10 meV and q = 5 A-1. While using thermal neutrons, the energy range is
10 times larger and the q range is about 2. 5 times larger. The spectrometer
resolution in both energy and q is typically a few percent but depends on the
collimation and monochromatization of the beam: higher resolution means less
flux and smaller count rate in the end. Modern spectrometers are flexible to
allow adapting the experimental conditions to the resolution requirements of
any particular experiment as much as possible. This flexibility is one of the
strongest points of triple axis machines.
Spectrometers which are built for the study of magnetic systems profit
very much from the ability to use polarized neutrons (half of all spectrometers
in the world either have this capability as an option, or use it all the time).
Polarization handling on a triple axis instrument is straightforward: one
typically uses permanent magnets creating a vertical guide field along the
spectrometer arms and an arrangement of Helmholtz coils for the sample
region to control magnitude and direction of the field at the sample position. In
real experimental conditions typical flipping ratios in the beam incident on the
sample (k; = 2 . 7, ... ,4 A-I) are in the range between 10 and 20. The option
to change the field direction with respect to q is important. As we have seen in
the theory Section 3. 1. 1, this has a crucial effect on how the magnetic
scattering changes the scattered beam polarization.
A particular example that we would like to highlight is the inelastic
100 G. Ehlers, F. Klose

scattering work performed on the germanate CuGe03' It shows the particular


value of polarization analysis in neutron scattering.
CuGe03 is an inorganic compound undergoing a "spin-Peierls" transition
below T Sp """ 14 K (Hase et al., 1993; Boucher and Regnault, 1996). This
means that linear chains of antiferromagnetically coupled Cu 2 + ions, which are
parallel to the c axis, show intrachain dimerization at low temperature: the
Cu 2 + ions are slightly displaced along the chain direction to form pairs of
nearest neighbors. This effect was predicted for one-dimensional Heisenberg
or XY antiferromagnetic s = 1/2 spin chains, where at low temperature large
quantum fluctuations give rise to a broad continuum of excitations (Muller
et al., 1981). These fluctuations induce the lattice distortion via strong spin-
phonon coupling. Experimentally, the spin-Peierls transition has been found by
measuring the magnetic susceptibility of a crystal in applied fields parallel to
the different crystal axes, and has been confirmed by specific heat, X-ray and
neutron diffraction, Raman scattering and electron spin resonance (ESR)
measurements. A third phase that appears in high magnetic fields above
12.5 T has been identified in which the whole lattice becomes incommensurate
(Kiryukhin et aI., 1996). Figure 3. 19 shows the (B, T) phase diagram of
CuGe03 (Boucher and Regnault, 1996).

18

16

14 U
12

EIO ,- Hllc
"v
(;: 8 .6. Hllb
Hila
6 Hila
D
4

0
4 6 8 10 12
Temperature (K)

Figure 3. 19 ( B, T) phase diagram of CuGe03 showing the three different phases:


uniform (U), dimerized (D) and incommensurate (I). Solid and dashed lines are second
and first order transitions, respectively. (L) is a Lifshitz point. (Boucher and Regnault,
1996).

The inelastic polarized neutron scattering work has significantly


contributed to our understanding of CuGe03 by measuring the structural and
magnetic excitations in the low-field phases. This has revealed some
Characterization of Magnetic Materials by Means of Neutron Scattering 101

peculiarities which are not accounted for by the standard theoretical approach.
For example, while in the dimerized phase there is an energy gap in the
spectrum of the magnetic excitations as expected (..1 SP :::::::; 2 meV ). this gap
does not vanish at the critical temperature T Sp (Regnault et al.. 1996). This
was attributed to the persistence of a "pseudo gap", broad inelastic scattering
present above T Sp and peaking at finite energy transfer. In the non-spin-flip
scattering (which should contain only nuclear scattering) the scattering
contribution at ..1 Sp is also visible, even after correction to the finite beam
polarization. This means that at this position in energy space the magnetically
and inelastically scattered neutron suffers an additional change of its
polarization, either a depolarization or a rotation around the sample field. As
we have seen in Section 3. 1. 3, three-dimensional polarization analysis can
not distinguish between these two cases.
An experiment using spherical polarization analysis was performed later to
elucidate this finding (Regnault et al.. 1999). It was found that some off-
diagonal elements of the polarization tensor, which can be determined by this
method, are non-zero at (q = (0, 1. 1/2), hw = 2meV), whereas no
polarization in the scattered beam appeared when the incident beam was
depolarized, which altogether was consistent with the earlier experiment,
proving a non-trivial rotation of the final beam polarization due to the
scattering. What does this mean? In the off-diagonal elements, terms appear
that couple nuclear and magnetic scattering amplitudes. In an inelastic
experiment these are correlation functions which couple the time dependences of the
spin components perpendicular to q and the time dependences of atomic
displacements parallel to q at the same site and the same time. Hence one can
qualitatively conclude that the low-energy excitations observed in CuGe03 are "dual
entities" with both spin and lattice degrees of freedom (Regnault et aI., 1999).

3.3.2 Detection of Slow Motions by Neutron Spin Echo


Neutron spin echo (NSE) is a technique for measuring the neutron energy
transfer in the scattering process with very high precision. Its main
applications are found in quasielastic scattering experiments (excitation
centered around zero energy transfer). Contrary to other inelastic techniques
such as TAS, it measures the correlation function in reciprocal space and time
I ( q, t). The time window spans from 10- 12 to 10 -7 s, partly overlapping with
the /JSR technique that can detect even slower spin dynamics but has no q
information. Compared to TAS, spin echo measures at longer times (smaller
energy transfer). Traditionally it is a domain of soft matter physics and
chemistry, and not many magnetic systems were studied in the past. Review
articles on the spin echo method are available. with emphasis on applications
in soft (Ewen and Richter, 1997) and condensed (Mezei, 1993) matter
physics. The most comprehensive reference, with many applications, is the
book-Neutron Spin Echo (mezei, 1979).
102 G. Ehlers, F. Klose

The functioning of a spin echo spectrometer can be described as follows


(Fig. 3.20). The beam is monochromatized to typically Ll/\//\ """ 15% by a
velocity selector. The basic idea of NSE is to compare the ingoing and
outgoing neutron velocities (before and after the scattering event,
respectively) by using the Larmor precession of the neutron spin in a magnetic
field (compare Section 3. 1. 2). The magnetic field is parallel to the beam
direction and is created in long solenoids. NSE uses a cold neutron beam that
is polarized by a supermirror transmission polarizer. After the polarizer, the
beam passes a Mezei spin flipper that acts as a TI/2 flipper: when the neutron
exits the flipper, its spin points perpendicular to the field direction. The
neutron now enters a region with a high magnetic field B I and on its way to the
sample the spin makes a total number of precessions given by

Velocity Spin rr/2 1 rr 2 rr/2 Spin Detector


selector polarizer Sample analyzer

"~m~ [] ~ C=:DO)D~ [] ~
Magnetic field profile ~
_t~~._
Spin precession uuuuuuuu . uuuuuuuu
Spin defocusing - ta::D t- cm
Figure 3.20 Layout of a generic NSE spectrometer. Top row: The beam enters on the
left side. The beam passes velocity selector, spin polarizer, n/2 flipper, precession field
8, , n flipper, sample, precession field 8 2 (note that the fields 8, and 8 2 are parallel),
n/2 flipper, spin analyzer, detector. The second row shows the corresponding magnetic
field profile. The third row demonstrates the manipulation of an individual neutron spin.
The fourth row schematically shows the defocusing of the spins due to the velocity spread
of the neutrons. At the first n/2 flipper, all spins point in the same direction, but they
refocus on the secondary side only if the product 8/ / v is the same on both sides (see
text). The sample (black dot) is centered between the precession fields 8, and 8 2 .

NI = 135/\ 65 X f
Bldl

where /\ is in A and the integral over the modulus of the field (in G cm) is to
be taken between the TI/2 flipper and the TI flipper. In modern spectrometers
the number N I can be as high as 2 x 10 4 Note that due to the wavelength
spread, N I is different for each individual neutron, so that the beam gets
dynamically depolarized after a short travel distance in the precession field.
All neutron spins are, however, in the plane perpendicular to the direction of
the magnetic field. The TI flipper turns the spin around an axis perpendicular to
the beam. This spin reversal effectively changes the sense of the spin
precession in the field B 2 , which is parallel to B I ' so that without energy
transfer due to sample scattering. if B I II = B 2 / 2 , the net precession angle of
Characterization of Magnetic Materials by Means of Neutron Scattering 103

each individual neutron at the second IT/2 flipper is zero, independent of the
actual number of precessions performed. Therefore at the second IT/2 flipper,
all spins refocus to the direction they had at the first IT/2 flipper and they are
finally flipped back to the original polarization direction. The quantity one
measures with the sample in place is the polarization of the beam at the
analyzer, which is the spectral average over all wavelengths
P NSE =P S <cos cp ( i\ ) )
of the total Larmor precession angle cp. The term coscp appears because the
IT/2 flipper turns only one of the precessing spin components back into the
beam direction. P s takes into account a possible change of the neutron
polarization by the scattering itself. In Section 3. 1. 3 we have seen that
nuclear spin incoherent scattering and magnetic scattering have this effect.
B 1 /, B2 /2
Under the condition ~ = ~ ,
VI V2
(v
1. 2 are the mean neutron velocities in

both spectrometer arms), it can be shown that

_fSeq, w)cos(tow)dw = I( q, t)
P NSE
- fseq, w)dw
I( q, 0)

yh BIll
with the correlation time ("Fourier time") t = - - - I(q, t) is called the
m vi'
intermediate scattering function.
The individual difference precession angles cp are in the first order
independent of the actual total number of precessions performed on both sides,
which is the reason why one can use a beam with a broad wavelength
distribution, hence with high intensity. It is one of the strong points of NSE that
energy resolution and q resolution are decoupled.
If the sample shows dynamics in the spin echo time window, P ( t) will
decrease as t increases,' because the sample scattering introduces a change
of the neutron velocity so that the spins do not perfectly refocus at the second
IT/2 flipper. However, one has to characterize the response of the
spectrometer to an elastic scatterer in order to distinguish between sample
dynamics and instrument imperfections which also decrease P ( t ).
Consequently, for each data set p~~nte ( q, t) a second data set PfJ~'rence ( q, t)
is required that is measured under identical experimental conditions using an
elastic reference sample. Dividing the two data sets p~~~le ( q, t) and
PfJ~'rence ( q, t) corresponds exactly to the usual deconvolution of the instrument
resolution in the energy domain (as done in triple axis spectroscopy) .
The spin dynamics in magnets are typically quite fast (in the ps range and
faster) and not easily observable by spin echo. Another difficulty is that
measuring real inelastic modes with spin echo as described here 1 -=I=- 2 ) (v v
requires a very complicated spectrometer setup. A way out of this problem is
provided by the possible combination of both TAS and NSE, where the spin
echo setup is put on top of a host triple axis spectrometer. This works
104 G. Ehlers, F. Klose

especially well with the resonance spin echo variant (Keller et al., 1998;
Koppe et al., 1999, references therein). In this technique the magnetic
precession fields and the rr/2 flippers are replaced by a pair of resonance spin
fl ippers on each spectrometer arm. In a resonance flipper one superposes a
high and homogeneous field 8 0 and a perpendicular radiofrequency field 8 1
cos(2rrVt). Typically, v~300 kHz and 80~ 100 G. In such a configuration a
rr flip is obtained, if the frequency v equals the Larmor frequency of the
neutron in the field 8 0 , and if the amplitude 8 1 of the oscillating field is set to
8 1 d = 135. 65 ; G(A~m (where d is the fl ipper thickness). The space
between the flippers is in strictly zero field, hence the spin keeps its direction
in the laboratory frame but precesses in the oscillating frame of the flippers.
An example that nicely demonstrates what neutron spin echo can do today
is provided by the recent work of Casalta and co-workers (Casalta et al.,
1999). They reported an experiment on the spin dynamics of mono-domain
iron clusters embedded in an insulating matrix of A1z03' With a mean Fe
particle diameter of 20A (standard deviation of 4 A) and a body-centered
cubic structure as shown by transmission electron microscopy (TEM), the
clusters behave as single-domain particles due to the lack of domain walls.
This leads to the existence of a "superspin" associated with the entire
particle. Inelastic neutron scattering showed the existence of two distinct
magnetic fluctuation components both associated with the whole spin of the
particle (superparamagnetic fluctuations). The faster relaxing component
could be resolved using triple-axis and time-of-flight spectroscopy and was
attributed to transverse fluctuations. The slow component was identified with
the longitudinal superspin fluctuations and could be resolved using the ultra-high
energy resolution of the spin echo technique (Fig. 3.21). The measurement at
a correlation time of 200 ns in this experiment is about equivalent to a
measurement at an energy transfer below 10 neV.
A hot topic in recent magnetism research is the study of "spin-ice"
H02 Ti 2 0 7 and related compounds. Here again, only the measurement of the
intermediate scattering function J (q, t) which contains all information on the
spin dynamics-spatial and temporal-has lead to a real understanding of the
magnetic interactions, when combined with what was known using other
techniques such as ac-susceptibility and j../SR. Spin ice H02 Ti 2 0 7 belongs to
the class of geometrically frustrated magnets. An extensive introduction into
the subject can be found in recent reviews (Ramirez, 2001; Bramwell and
Gingras, 2001).
The magnetic Ho3+ ions occupy a cubic pyrochlore lattice (space group
Fd3m) of corner-I inked tetrahedra (Fig. 3. 22). The magnetic moments are
constrained by the crystal electric field (CEF) to local (111) axes. This
frustrates the dominant (effectively ferromagnetic) dipolar interactions in the
system and leads to frozen, non-collinear, spin disorder below -1 K. The
spin ice state is in a way analogous to the Pauling hydrogen disorder of water
ice (H 2 0), with each spin equivalent to a hydrogen displacement vector
situated on the mid-point of an oxygen-oxygen Iine of contact, hence the name
Characterization of Magnetic Materials by Means of Neutron Scattering 105

0.1 I 10 100
Time (ns)

Figure 3. 21 Normalized intermediate scattering function at q 0.07 A-1 for various


temperatures. Below 200 K a specially derived model function is used to fit the data
(higher temperature: single exponential relaxation) taking into account a distribution of
single relaxation times due to the spread in particle size (Casalta et ai., 1999).

spin ice (Harris et al., 1997). The single ion ground state is an almost pure
I J, M J > I = 18, 8> doublet with < 111 >quantization axis, separated by over
200 K from the first excited state (Rosenkranz et al., 2000).
Fig. 3.23 shows the normalized relaxation function F(q, t) = /(q, t)j
/(q,O) measured for a polycrystalline sample of Ho2 Ti 2 0 7 at the IN11
spectrometer (ILL). At all temperatures between O. 05 K and 200 K the

Figure 3. 22 The pryrochlore lattice. In cubic pyrochlores of chemical composition


A, B, 0 7 both A and B atom sublattices independently form this network of corner-sharing
tetrahedra.
106 G. Ehlers, F. Klose

relaxation can be fitted with excellent precision to a simple exponential


function
F(q, t) A exp {- V ( T) t} ,

1.0

0.8

0 0.6
c;.
'<
Q
c;. 0.4
'<

0.2

0
4 6 8 4 6 8 4 6 8
0.01 0.1
Correlation time (ns)
(a)

1.0

0.8 !; T=IOO K

0.6
tt t! q=0.7 k l
0
c;.
'<
Q 0.4
tt. ! o q=1.1 k
.. q=1.4 k
l
l

c;.
'< t,

0.2

4 6 8 4 6 8 4 6 8
0.01 0.1
Correlation time (ns)
(b)

Figure 3. 23 The intermediate scattering function measured for a polycrystalline sample of


H02 Ti 2 0, at the IN11 spectrometer (ILL). Top panel: at different temperatures showing
the speed-up when temperature is increased. Lower panel: at different q' s showing the
q-independence of the dynamics.

where A = O. 91 O. 01. The frequency v ( T) can be fitted to an Arrhenius


expression v (T) = 2 Va exp( - E aj k B T) with attempt frequency Va = (1. 1 X
1011 0.2 X 1011) Hz and activation energy E a = (293 12) K. The origin of
the dynamics observed in NSE can be identified by the experimental findings:
the activation energy is close to the first group of CEF levels (Rosenkranz et
ai., 2000), and the q independence of the scattering indicates negligible two-
Characterization of Magnetic Materials by Means of Neutron Scattering 107

spin correlations, so that the dynamics must be due to a single ion spin flip
mechanism between the two states of the ground state doublet.
By Fourier transform, the NSE signal can be related to the imaginary part
of the generalized susceptibility X" (q, w) which can be extrapolated to v <
10 5 Hz for comparison with the bulk ac-susceptibility. This analysis gives
vZ ( T) iwv ( T) }
X(q, w) = X(q) { VZ(T) + w Z + VZ(T) + wZ

where k B Thw was assumed. X(q= 0) can be well approximated as -liT


in the relevant temperature range. Inserting v (T) as above one finds that in
the ac susceptibility a peak should be observed around 15 K. Such a peak was
indeed found (Ehlers et ai., 2002) on a single crystal sample of HozTi z0 7 in a
finite field. In fact, when X" was normalized to X(O), it could be seen even in
very low field. Its frequency shift implies an activation barrier of -250 K, as
expected.
Towards low temperature, the susceptib iI ity was found to increase
significantly, though, showing a second peak at - 1 K peak associated with an
activation energy Ea":::::::: 20 K and attempt frequency v; - 10 10 Hz. This should
correspond to a second, even slower dynamical process. It is limited to low
temperature (below -15 K) and low frequency, < 105 Hz (invisible at least in
the frequency domain of NSE), and approximately independent of temperature
above -4 K. The thermal activation energy of this second process is closer to
the other major energy scale in the system: the dipolar interaction, which was
estimated to be 2.4 K for spin ice HozTi z 0 7 (Bramwell and Gingras, 2001).
Below T - 15 K, where the spins freeze, they are subject to an unusually
strong, slowly fluctuating, dipolar field, created by the nearest neighbors. In a
classical picture, near neighbor spins are fixed along axes at 109. 5 to each
other, and so experience a mutual torque that cannot be eliminated by local spin
reorientation. To explain the very slow dynamics below T-15 K, it was therefore
suggested that in the temperature range of 4 to 15 K the transverse component of the
dipolar field mixes higher IMJ >states into the ground state, causing the single ion
ground state to no longer be an almost pure I 8> doublet, and inducing a finite rate
of spin inversion that is temperature-independent. This new "quantum relaxation
channel" is only possible in the paramagnetic phase above 1 K where the mean
dipolar field is zero.

3.4 Summaries

Due to its unique elementary characteristics, the neutron is especially suited


for probing the magnetic properties of materials. It is not electrically charged,
and therefore penetrates deeply into condensed matter. On the other hand, it
possesses spin, and thus interacts with atomic nuclei as well as with magnetic
moments present in matter. These two types of interactions are of comparable
108 G. Ehlers, F. Klose

strength. Neutrons with wavelengths in the A range possess kinetic energies in


the meV range, which are the typical energies of elementary excitations in
condensed matter. Consequently, neutrons can simultaneously probe structural
and magnetic spatial correlations on atomic to mesoscopic length scales, as
well as structural and magnetic temporal correlations in the range of 10 - 14 to
10- 7 S. By keeping track of the neutron spin orientation and its change during
the interaction with the sample, one can unambiguously separate nuclear and
magnetic scattering processes.
Compared to other experimental techniques developed to investigate
magnetic properties of matter, neutron scattering has particular merits:
CD transparent and easy experimental procedures that allow a straightforward
conversion of experimental data into physical quantities, ~ it is a non-
destructive technique sensitive to both volume and surface properties, and
@ it is possible to study a huge variety of magnetic phenomena and different
classes of materials. The dependence of magnetic properties on temperature,
pressure, or magnetic field can easily be explored since most sample
environments can be made transparent to neutrons.
We have presented basic instrumental concepts for scattering experiments
using spin polarized neutrons, focusing on appl ications in condensed matter
magnetism research. To summarize, an overview on different particular
applications for the different techniques is given in the Table 3. 1.

Table 3 1

Technique Applications

Powder and single Ordering pallerns of magnetic moments,


crystal diffraction magnetization density maps

Condensed mailer: atomic or magnetic


Diffuse scallering
short range order
Elastic experiments Magnetization depth profile in thin films,
(Structural Information) Reflectometry orientation and lateral arrangements of
magnetization vectors
Investigations on mesoscopic length scales
Small angle scallering
(10 A to 10000 A), such as grains or
(SANS)
magnetic domains
Triple axis and chop-
Inelastic experiments per spectroscopy Magnetic excitations (spin waves)
(dynamical information)
Neutron spin echo Slow dynamics (e. g., in disordered magnets)

References
Als Nielsen, J. In: Topics in Current Physics: Structure and Dynamics of
Surfaces. Springer, Berlin(1986)
Characterization of Magnetic Materials by Means of Neutron Scattering 109

Ankner, J. F. and G. P. Felcher. J. Magn. Magn. Mater. 200: 741 (1999)


Baibich, M. N., J. M. Broto, A. Fert, F. Nguyen van Dau, F. Petroff, P.
Etienne, G. Creuzet, A. Friedrich and J. Chazelas. Phys. Rev. Lett. 61:
2472 (1988)
Boucher, J. P. and L. P. Regnault. J. Phys. I France 6: 1939 (1996)
Bramwell, S. T. and M. J. P. Gingras. Science 294: 1495 (2001)
Brown, P. J. Physica B 297; 198 (2001)
Brown, P. J., J. B. Forsyth and F. Tasset. Physica B 267-268:215 (1999)
Brown, P. J. In: A. J. C. Wilson, ed. International Tables for
Crystallography, Vol. C, chapter 4. 4. Kluwer Academic Publishers,
Dordrecht, p. 391, 1992
Cable, J. W., S. A. Werner, G. P. Felcher and N. Wakabayashi. Phys.
Rev. Lett. 49:829 (1982)
Casalta, H., P. Schleger, C. Bellouard, M. Hennion, I. Mirebeau, G.
Ehlers, B. Farago, J. - L. Dormann, M. Kelsch, M. Linde and F. Phillipp.
Phys. Rev. Lett. 82: 1301 (1999)
Coulter K. P., T. E. Chupp, A. B. McDonald, C. D. Bowman, J. D. Bowman,
J. J. Szymanski, V. Yuan, G. D. Cates, D. R. Benton and E. D. Earle.
Nucl. Instr and Meth. A 288: 463 (1990)
Cussen, L. D., D. J. Goossens and T. J. Hicks. Nucl. Instr. Meth. Phys.
Res. A 440:409 (2000)
Cywinski, R., S. H. Kilcoyne and J. R. Stewart. Physica B 267 - 268: 106
(1999)
Desclaux, J. P. and A. J. Freeman. In: A. J. Freeman and G. H. Lander,
eds. Handbook on the Physics and Chemistry of the Actinides, Vol. 1. North
Holland, Amsterdam, p. 46 ( 1984)
O'Donovan, K. V., J. A. Borchers, C. F. Majkrzak, O. Hellwig and E. E.
Fullerton. Phys. Rev. Lett. 88:067,201 (2002)
Ehlers, G., A. L. Cornelius, M. Orend ac, M. Kajcakova, T. Fennell, S. T.
Bramwell and J. S. Gardner. J. Phys. C Condensed Matter (2002)
(submitted for publication)
Eilhard, J., A. Zirkel, W. TschOp, O. Hahn, K. Kremer, O. Scharpf, D.
RichterandU. Buchenau.J. Chem. Phys. 110:1819 (1999)
Ewen, B. and D. Richter. Neutron Spin Echo Investigations on the Segmental
Dynamics of Polymers in Melts, Networks and Solutions. In: Advances in
Polymer Science, Vol. 134. Springer Verlag, Berlin, Heidelberg( 1997)
Felcher, G. P. and S. G. E. te Velthuis. Appl. Surf. Sci. 182: 209 (2001)
Freund, A., R. Pynn, W. G. Stirling and C. M. E. Zeyen. Physica 120 B: 86
(1983)
Fullerton, E. E., J. S. Jiang and S. D. Bader. J. Mag. Mag. Mater. 200: 392
( 1999)
Furrer, A. (ed.), Magnetic Neutron Scattering. In: Proc. of the Third
Summer School on Neutron Scattering. World Scientific, Singapore( 1995)
Gabrys, B. J. Physica B 267-268: 122 (1999)
110 G. Ehlers, F. Klose

Garcia-Hernandez, M., F. J. Mompean, O. Scharpf, K. H. Andersen and B.


Fak. Phys. Rev. B 59 :958 (1999)
de Gennes, P. G. Theory of Neutron Scattering by Magnetic Crystals. In: G.
T. Rado, H. Suhl eds. Magnetism Vol. 3, Academic Press, New York
(1963)
Groom, D. E. et al. The European Physical Journal C 15: 1 (2000)
Grunberg, P., R. Schreiber, Y. Pang, M. B. Brodsky and H. Sowers. Phys.
Rev. Lett. 57: 2442 (1986)
Halpern, O. and M. H. Johnson. Phys. Rev. 55: 898 (1939)
Harris, M. J., S. T. Bramwell, D. F. McMorrow, T. Zeiske and K. W.
Godfrey. Phys. Rev. Lett. 79: 2554 (1997)
Hase, M., I. Terasaki and K. Uchinokura, Phys. Rev. Lett. 70:3651 (1993)
Heil, W., J. Dreyer, D. Hofmann, H. Humblot, E. Lelievre-Berna and F.
Tasset. Physica B 267 - 268: 328 (1999)
Hiess, A., P. J. Brown, E. Lei ievre-Berna, B. Roessl i, N. Bernhoeft, G. H.
Lander, N. Aso and N. K. Sato. Phys. Rev. B 64: 134,413 (2001)
Hjorvarsson, B., J. A. Dura, P. Isberg, T. Watanabe, T. J. Udovic, G.
Andersson and C. F. Majkrzak. Phys. Rev. Lett. 79: 90 1 (1997)
Hoffmann, A., M. R. Fitzsimmons, J. A. Dura and C. F. Majkrzak. Phys.
Rev. B 65:024,428 (2002)
Kasuya, T. Prog. Theor. Phys. 16:45 (1956)
Keller, T., R. Golub, F. Mezei, R. Gahler. PhysicaB241-243:101 (1998)
Keller, T., T. Krist, A. Danzig, U. Keiderling, F. Mezei and A.
Wiedenmann. Nucl. Instr. Meth. Phys. Res. A 451 :474 (2000)
Kiryukhin, V., B. Keimer, J. P. Hill and A. Vigliante. Phys. Rev. Lett. 76:
4608 (1996)
Klose, F., Ch. Rehm, D. Nagengast, H. Maletta and A. Weidinger. Phys.
Rev. Lett. 78: 1150 (1997)
Koppe, M., M. Bleuel, R. Gahler, R. Golub, P. Hank, T. Keller, S.
Longeville, U. Rauch and J. Wuttke. Physica B 266: 75 (1999)
Krist, T., F. Klose and G. P. Felcher. Physica B 248:372 (1998)
Lamelas, F.J., S. A. Werner, S. M. Shapiro and J. A. Mydosh. Phys. Rev.
B 51 :621 (1995)
Lee, W. T and F. Klose. Neutron Optics. In: J. L. Wood, I. S. Anderson eds.
Proceedings of SPIE Vol. 4509p. 145 (2001)
Lee, W. T., F. Klose and B. P. Toperverg. Physica B(2002a)(submitted for
publication)
Lee, W. T., S. G. E. te Velthuis, G. P. Felcher, F. Klose, T. Gredig and
D. Dahlberg. Phys. Rev.B65:224, 417 (2002b)
Lovesey, S. W. Theory of Neutron Scattering from Condensed Matter.
Clarendon Press, Oxford( 1984)
Majkrzak, C. F. Physica B 173: 75 (1991)
Meneghetti, D. and S. S. Sidhu. Phys. Rev. 105:130 (1957)
Mezei, F. (ed.). Neutron Spin Echo. In: Lecture Notes in Physics, Vol.
Characterization of Magnetic Materials by Means of Neutron Scattering 111

128. Springer Verlag, Berlin, Heidelberg(1979)


Mezei, F. Liquids, Freezing and Glass Transition. J. P. Hansen, D. Levesque
and J. Zinn-Justin eds. Les Houches, Session LJ, 1989. Elsevier( 1991)
Mezei, F. Int. J. Mod. Phys B 7:2885 (1993)
Mezei, F. and P. A. Dagleish.Comm. Phys. 2:41 (1977)
Mezei, F., G. Ehlers, C. Pappas, M. Russina, T. J. Hicks and M. F. Ling.
Physica B 276-278:543 (2000)
Moon, R. M., T. Riste and W. C. Koehler. Phys. Rev. 181 :920 (1969)
Mueller-Buschbaum, P., J. S. Gutmann, R. Cubitt, M. Stamm. Colloid
Polym. Sci.277:1193 (1999)
Muller, G., H. Thomas, H. Beck and J. C. Bonner. Phys. Rev. B 24: 1429
( 1981)
Mughabghab, S. F., M. Divadeenam, N. E. Holden. In: Neutron Cross
Sections, Vol. 1. Academic Press, New Yorke 1981)
Murani, A. P., O. Scharpf, K. H. Andersen, D. Richard and R. Raphel.
Physica B 267-268: 131 (1999)
Parratt, L.G. Phys. Rev. 95:359 (1954)
PasseII , L. and R. I. Schermer. Phys. Rev. 150: 146 (1966)
Prinz, G. A. J. Mag. Mag. Mater. 200 :57 (1999)
Ramirez, A. P. In: K. H. J. Buschow, ed. Handbook of Magnetic Materials
Vol. 13, Chapter 4, Elsevier, Amsterdam(200 1)
Regnault, L. P., M. Aln, B. Hennion, G. Dhalenne and A. Revcolevschi.
Phys. Rev. B 53:5579 (1996)
Regnault, L. P., F. Tasset, J. E. Lorenzo, T. Roberts, G. Dhalenne and A.
Revcolevschi. Physica B 267 - 268: 227 (1999)
Rehm, Ch., F. Klose, D. Nagengast, H. Maletta and A. Weidinger.
Europhys. Lett. 38: 61 (1997)
Ressouche, E., J. X. Boucherle, B. Gillon, P. Rey and J. Schweitzer. J.
Am. Chem. Soc. 115:3610 (1993)
Rich, D. R., S. Fan, T. R. Gentile, D. Hussey, G. L. Jones, B. Neff, W.
M. Snow and A. K. Thompson. Physica B 305:203 (2001)
Rosenkranz, S., A. P. Ramirez, A. Hayashi, R. J. Cava, R. Siddharthan
and B. S. Shastry. J. Appl. Phys. 87: 5914 (2000)
Rossat-Mignod, J. Neutron Scattering. In: K. Sk6ld, D. L. Price eds.
Methods of Experimental Physics, Vol. 23. Academic Press, London( 1987)
Ruderman, M.A. and C. Kittel. Phys. Rev. 96:99 (1954)
Scharpf, O. and H. Capellmann. Phys. Stat. Sol. A 135:359 (1993)
Schreyer, A., C. F. Majkrzak, N. F. Berk, H. Grull and C. C. Han. J. Phys.
Chem. Solids 60: 1045 (1999)
Shirane, G., S. M. Shapiro and J. M. Tranquada. Neutron Scattering with a
Triple-Axis Spectrometer. Cambridge University Press, Cambridge(2002)
Speriosu, V.S., D.A. Herman, Jr., I.L. Sanders and T. Yohi. IBM J. Res.
Develop. 44: 186 (2000)
Squires, G. L. Thermal Neutron Scattering. Cambridge University Press,
112 G. Ehlers, F. Klose

Cambridge( 1978)
Stewart, J. R., K. H. Andersen, R. Cywinski and A. P. Murani. J. Appl.
Phys 87: 5425 (2000)
Tasset, F., P.J. Brown, E. Lelievre-Berna, T. Roberts, S. Pujol, J. Allibon
and E. Bourgeat-Lami. Physica B 268: 69 (1999)
Temst K., M. J. Van Bael and H. Fritzsche. Appl. Phys. Lett. 79:991 (2001)
Thompson, D. A. and J. S. Best. IBM J. Res. Develop. 44: 311 (2000)
Toperverg, B. P., G. P. Felcher, V. V. Metlushko, V. Leiner, R. Siebrecht
and O. Nikonov. Physica B 283: 149 (2000)
Toperverg, B. P. Physica B 297: 160 (2001)
Weissmuller, J., A. Michels, J.G. Barker, A. Wiedenmann, U. Erb and R.
D. Shull. Phys. Rev. B 63:214,414 (2001)
Wiedenmann, A. Physica B 297: 226 (2001)
Williams, W. G. Polarized Neutrons. Clarendon Press, Oxford( 1988)
Windsor, C. G. Pulsed Neutron Scattering. Taylor & Francis Ltd., London and
Halsted Press, New York ( 1981)
Yosida, K. Phys. Rev 106:893 (1957)
Zabel, H. Physica B 198: 156 (1994a)
Zabel, H. Appl. Phys. A-Mater. 58: 159 (1994b)

We are grateful to Ch. Rehm, W. T. Lee, C. E. Prokuski, A. Hoffmann, S. G. E. te


Velthuis,G. P. Felcher, P. Bani, K. H. Andersen, R. S. Stewart, A. P. Murani and E.
Lelievre-Berna for their contributions to this article and for enlightening discussions. This work
is supported by the Spallation Neutron Source Project (SNS). SNS is managed by
UT-Battelle, LLC, under contract No. DE-AC05-000R22725 for the U. S. Department of
Energy.
4 Advanced Transmission Electron Microscopy of
Nanostructured Magnetic Materials

Y. Liu, J. Sellmyer

4. 1 Introduction

Transmission electron microscopy has become the most powerful tool for
characterizing the structure of magnetic materials. We can divide the structure
of magnetic materials into physical structure such as crystal structure,
morphology, grain size, texture, grain boundaries, volume fraction of each
phase, inter-phase interfaces, surfaces, defects, etc. and magnetic structure
such as magnetic-domain structure, and recording pattern in magnetic
recording media. This chapter wi II focus on the physical structure of magnetic
materials. The advantages using TEM include: Sensitivity: structural
properties for a few atomic layers in a film can be readily examined by plan-
view or cross-section samples. Resolution: TEM is the only tool that can
provide the internal structure information of a material at the atomic level. The
reported resolution achievable by TEM exceeds O. 1 nm. The object that can
be analyzed is also extremely small: as small as a single heavy atom.
Simplicity: in the imaging mode when the conditions are set right the image is
simply the projection of the structure. This eliminates the tedious image or
diffraction simulation. Versatility: TEM can be used in both image mode and
diffraction mode or combined image mode and diffraction mode. Therefore
both structural and morphological information can be retrieved. Compositional
analysis by X-ray energy dispersive spectrometry or energy loss spectroscopy
are conveniently integrated in commercial TEMs. In general, all conventional
techniques can be applied to magnetic materials with attention paid to the
following: first there is a strong interaction between the magnetic specimen
and the objective lens. This can cause problems as the specimen may jump out
of the holder and fall into the column of the TEM. Associated with this
interaction is the astigmatism correction. As the magnetic field of the
specimen interferes with the objective lens stigmation is needed for HRTEM
whenever the specimen is moved. Another issue is that almost all the
advanced magnetic materials require the optimum nanostructure design.
Therefore, nanostructure characterization is a predominant part of magnetic
materials. This chapter reviews important TEM techniques and their
114 Y. Liu, J. Sellmyer

applications to magnetic materials characterization. First specimen-


preparation methods for magnetic samples are reviewed. Advanced TEM
techniques with emphasis on nanostructured magnetic materials are
elucidated. Examples are given to demonstrate the effectiveness of each
technique.

4.2 Specimen Preparation

All the traditional specimen preparation techniques such as electro-polishing,


ion milling etc. can be applied to magnetic sample preparation. A thorough
review is given by Robert et al. (2002). The major challenge for preparing
magnetic samples is the brittleness of most magnetic samples. Here we
review several methods with emphasis on special magnetic samples.

4. 2. 1 Bulk Samples

Most magnetic materials have complicated crystal structures which make them
very brittle. The magnetic force between the objective lens and the specimen
(the interference of the specimen to the magnetic field of the objective lens) is
proportional to the volume of the specimen. Therefore it is very desirable to
reduce the volume of the magnetic sample. One method is to embed the
magnetic material with epoxy in a non-magnetic tube such as a Cu tube of
3 mm diameter. The specimen is then cut into 0.2 mm slices. The embed
epoxy can be purchased from commercial vendors such as Electron Microscopy
Sciences. Mechanical polishing, dimpling and ion milling can be applied to
make the final sample. This means the sample has much reduced volume, and
is less prone to fracture.

4. 2. 2 Magnetic Thin Films

Magnetic thin films are usually deposited on glass, Si, quartz or AI alloy. In
many cases special texture may develop in thin films. Frequently both plan-
view and cross-sectional magnetic thin film samples are needed to retrieve
important structural information. For plan-view samples the thickness of the
sample is measured at each thinning step. The thinning steps involve
mechanical dimpling, and ion milling. The standard procedure is: flat wheel
dimpling to 200 IJm , convex wheel dimpling to 5 IJm, ion milling at the
substrate side at 6 kV until a hole is opened. A final thinning by ion mill at 2 kV
for 5 min is used to eliminate/reduce the amorphous layer. The cross-sectional
Advanced Transmission Electron Microscopy of Nanostructured. . . 115

sample is made of many layers of the specimen embedded in a metal tube


(usually Cu tube). Standard mechanical polishing, dimpling and ion milling are
then applied to make the final sample as for the case of plan-view samples.
In the following we demonstrate an example to use a combination of cross-
section sample and plane-view sample to reveal the nanostructure in Au-Cu
multilayer structure and provide an explanation to the observed magnetic
properties. Artificial thin films consisting of periodic stacking layers of a
magnetic metal and a non-magnetic metal such as multilayer ColM (M = Pt or
Pd, Au, Cu, Ag, Ir) are attracting strong interest (Broeder et ai., 1991).
Such films could have in-plane anisotropy or perpendicular anisotropy
depending on the thickness of the magnetic layer. The anisotropy of the films
changes from in-film plane to perpendicular direction as the thickness of the Co
layer reduces to a critical value (Carcia, 1988; Chappert, 1988). The films of
perpendicular magnetic anisotropy and significant Kerr rotation have potentials
as high-density magneto-optic storage devices. The anisotropy of the films is
influenced by different processing routes such as molecular beam epitaxy,
electron beam evaporation, and sputtering. Defect structures such as the
roughness of the interface (Bruno, 1988) and the degree of mixing of the
atoms at the interface (the sharpness of the interface) have been found to
affect the anisotropy (Broeder, et al., 1988). The role of the non-magnetic
layer is to create interfaces as many as possible. For this reason thinner Au
layer is preferable. However, as the Au layer becomes thinner, the periodic
multilayer structure could become unstable and the perpendicular anisotropy of
the film may be destroyed. The Au-Co multilayer films were deposited by DC
magnetron sputtering at room temperature. First, about 40 nm of Au was
deposited on a Si substrate. The Co layer thickness was set at O. 5 nm and the
Au layer thickness varies from O. 5 to 8 nm. Figure 4. 1 shows the
magnetization loops for two films (Au60 A/C05 A) x 91 Au 400 A (denoted
Au60) and (Au 10 AI C05 A) x 501 Au400 A (denoted Au 10). The fi 1m Au60
shows a well defined perpendicular magnetization loop with loop squareness
S = 1 while the film Aul0 show an irregular perpendicular magnetization loop
with S = O. The measured magnetic anisotropy is 3 x 106 ergl cm 3 for Au60 and
1. 5 x 106 ergl cm 3 for Au 10 .
The effort of TEM work was directed toward the question: what is the
cause of the different values of the anisotropy of the two samples Au60 and
Aul0. Both plane-view samples and cross sectional TEM samples were
prepared as described earlier. Figure 4. 2 shows the TEM bright field
micrograph of the Au60 film and Au 10 film. The grain size of the Au underlayer
is about 50 nm. Slight curvature of the multilayer is observed. The curvature is
inherited from the surface of the underlayer. Both nanodiffraction and HREM
show that the multilayer is epitaxially grown on the Au underlayer. Figure 4. 3
shows a [110J zone axis high resolution electron microscopy (HREM) image of
the Au60 film. The Au layers and the Co layers can be identified by their
different stacking mode. The Au layers have the ABCABC stacking with
116 Y. Liu, J. Sellmyer

-10 o +10 -10 o +10


H(kOe) H(kOe)
(a) (b)

Figure 4.1 Magnetization loop of (a) AulO and (b) Au60 films.

frequent twins . The letter T indicates a twin. The two atomic Co layers
interrupts the ABCABC stacking of Au and could be fitted to the AB stacking as
indicated by the arrows. The formation mechanism of the curvature of the
multilayer is clearly revealed. That is, the curvature is formed by many atomic
steps rather than the bending of the atomic planes.

(a) (b)

Figure 4. 2 Bright field TEM image of (a) Au60 film and (b) AulO film.

In the AulO film as seen in Fig. 4. 2b, the Au layers and the Co layers
can be differentiated from each other by their different contrast. The Au layers
appear dark because of their larger scattering factor to electrons. A careful
examination indicates that the multilayer structure is frequently interrupted by
local clustering which appears darker or brighter as indicated by the arrows.
The cluster size is in the range of 2 to 5 nm. Figure 4. 4 is an example of many
[llOJ zone axis HREM images taken from the AulO sample. The Au layers and
the Co layers in the HREM image cannot be differentiated. Local ABC stacking
(fcc structure), which is Iikely to be Au atom clusters, is observed. The
close-packed atomic planes are resolved in most of the place. However, the
majority part of the image does not reveal a clear ABCABC stacking. 1/3
Advanced Transmission Electron Microscopy of Nanostructured... 117

Figure 4. 3 [1 lO]zone axis HREM image of the Au60 film.

[lll]a dislocations are frequently observed as indicated by the arrow. The


appearance of such dislocations is consistent with the lattice mismatch
between Au and Co.

Figure 4.4 [1 lO]zone axis HREM image of the AulO film


118 Y. Liu, J. Sellmyer

The reader may ask the question why the black contrast in Fig. 4. 2 does
not necessarily represent Au cluster. It could be due to other contrast
mechanisms such as diffraction. The plan-view image in this case provides a
more convincing argument as shown in Fig. 4. 5. The white clusters are Co and
darker clusters are Au as confirmed by EDX analysis. A Au cluster is
surrounded by a donut-like Co cluster. Single isolated Co clusters are also
observed but in three dimension Au clusters must have been formed near them
as the extra Au atoms are squeezed out from the layer structure. The formation
of such Co and Au cluster is driven by the reduction of the interfaces (Liu
et aI., 1997).

Figure 4.5 Plane-View image of Au10 film. The clusters are indicated.

4. 2. 3 Magnetic Nanowires
A great motivation for magnetic materials research has been the enhancement
of recording density in computer rigid disks. Self-assembled nano-pores in
alumina serve as an template for embedding nanowires of various magnetic
materials. Plan-view and cross-sectional samples can be prepared by standard
techniques as described above. Another handy way to release the wires is to
simply dissolve the alumina using O. 2MH 2 Cr04 /0. 4MH 3 P0 4 solution (Liu
et al., 2002). The chemical reaction speed can be controlled by adjusting the
temperature slightly higher than room temperature. The alumina is etched
away and the freed wires appear dark tiny object on the surface of AI foil. The
released wires can be picked up by metal grid coated with a carbon film.
Figure 4. 6a is an image of the template, Fig. 4.6b is released Ni nanowires
and Fig. 4.6c is Fe/Pt multilayer nanowires. References for properties of
these nanowires can be found in Zeng et al., (2000).
Advanced Transmission Electron Microscopy of Nanostructured... 119

(a) (b)

10 nl11

Pt

5nl11

(c)

Figure 4.6 (a) plane-view of nano pore template, and (b) Ni nanowires released, and
(c) HRTEM image of Fe/Ptnanowires.

4.2.4 Magnetic Powders


For nonmagnetic powders such as Pt it is possible to dissolve them in acetone
and then pick them up with metal grid coated with C. This method is useful for
small magnetic powders with a grain size range smaller than 20 nm but not
desirable for magnetic particles larger than 100 nm as the interaction with the
objective lens may simply cause the particles to fly to the pole piece of the
objective lens. An easy way to avoid this is to mix the powder with epoxy and
then embed the mixture in a metal tube for curing. Standard sectioning,
polishing, dimpling, and ion milling are applied to make the sample for TEM.
One interesting way of preparing magnetic samples is to make use of the
magnetic field for finding certain crystallographic axis. For example, in the
120 Y. Liu, J. Sellmyer

C0 17 8m2 magnet the c-axis is the magnetic easy axis. The powder of C0178m2
is al igned in a magnetic filed and then sintered or cured in epoxy. TEM
samples can be made either having the c-axis parallel to or perpendicular to
the foil plane. The Zr rich precipitate is only observable in the specimen 8m2
(CoFeCuZr) 17 with the c-axis parallel to the foil plane as shown in Fig. 4.7.
Finding such precipitate in a randomly oriented sample requires skill of tilting
the sample to the preferred orientation.

(a) (b)

Figure 4. 7 Microstructure of 8m2 (CoFeCuZr) 17 magnet (a) c-axis perpendicular to


image plane, (b) c-axis parallel to image plane.

4. 2. 5 Special Techniques
Electro-polishing and ion-milling are the most popular tools for TEM specimen
preparation. Other methods for magnetic samples such as ultramicrotomy and
focused ion beam method are useful for special samples. Ultramicrotomy is the
mature technique for preparing biological samples. This technique, if used
properly, can be efficient for preparation of magnetic powders or thin films.
Usually the preparation of a cross-section sample for thin films is a time-
consuming process. If the thin film can be removed from the substrate and
embedded in epoxy, cross-section samples can be readily sliced by the
ultramicrotomy technique. Magnetic powders can also be embedded in epoxy
and then sliced by the microtome. Ultramicrotomy is a very productive process
and many slices ready to examine can be made in an hour. The latest
generation of focused ion beam (FIB) instrument can generate a programmable
nanoprobe to just thin the area of a few microns for TEM observation. Use of
FIB not only significantly enhances the efficiency but also provide very thin
Advanced Transmission Electron Microscopy of Nanostructured. . . 121

samples. A session was devoted to specimen preparation by FIB in the recent


Microscopy and Microanalysis Conference (Voelkl, 2002).

4. 3 Electron Diffraction

Electron diffraction has been widely used to solve materials problems. There
are three techniques for electron diffraction: selected area diffraction (SAD),
convergent beam electron diffraction (CBED) and nanodiffraction.

4. 3. 1 SAD Pattern and the Ring Pattern


Traditional SAD: The area in the specimen contributing to the diffraction
pattern is defined by using an aperture of desired size. Such apertures can
select an area in the specimen from 1 to 150 IJm. Figure 4. 8 shows the ray
diagram (a) for traditional SAD (b) using convergent angle to define the
selected area. In materials with grain size larger than 1 IJm, the SAD pattern is
a spot pattern produced by a single grain. A diffraction pattern map is made of
many diffraction patterns of major zone axes in the stereo projection. The
diffraction pattern map method has been widely used for phase identification
(Liu and Mazumder, 1994).

Aperture

Diffraction pattern Diffraction pattern


(a) (b)

Figure 4.8 Ray diagram (a) traditional SAD, (b) using convergent angle to define the
selected area.

The ring pattern: In a material with grain size around 10 nm, the SAD
pattern becomes a ring pattern produced by many grains. Such ring pattern is
most useful for phase identification of nanostructured materials. The key steps
122 Y. Liu, J. Sellmyer

for getting a ring pattern are:


( 1) Set magnification at the right value 40, 000 for JEOL 2010, for
example. At this magnification it is easier to locate the selected area
aperture.
(2) Set the objective current at a preset value. This is necessary to
ensure a cal ibrated camera length to be used for all diffraction patterns.
(3) Focus the image using z-axis control. This ensures that both the
height of the specimen and the objective lens is set at the same values for all
diffraction patterns.
(4) Insert the selected area aperture.
(5) Set condenser lens at maximum value. This setting provides a
parallel beam.
(6) Go to diffraction mode and choose proper camera length.
(7) Focus the spot pattern and make the exposure.
Using convergence of the beam to define the selected area: A ring pattern
is an efficient way for quick phase identification as the position of the ring is
well defined and not affected by thickness. It is desirable to collect such a ring
pattern for larger grain size material or from particles which usually do not form
a ring because of insufficient number of grains. An effective way to collect a
ring pattern is to adjust the convergence angle using the brightness knob. The
ray diagram is shown in Fig. 4.8b. The illumination area can be controlled at
quite large rang. The diffraction pattern is focused using the focusing lens. The
camera length for such patterns needs calibration under the same condition.
Figure 4. 9 shows an example of phase identification of two important
magnetic phase NdFe14 B2 and C05 Sm using such ring patterns. The calculated
intensity is deduced from the structure factor of each reflection. If there is no
texture in the material which is true in this case, the matches between
calculation and experiment are excellent. It needs to be mentioned that such
matching is qualitative when the specimen is thick. It has better accuracy for
thin crystals, suitable for nano-materials with a grain size less than 10 nm.

4.3.2 Convergent Beam Electron Diffraction


Convergent beam electron diffraction (CBED) has become a standard
technique for many applications. Following lists the important ones: Phase
identification (Mansfield, 1989), point group and space group identification
(Buxton et ai., 1976; Liu et al., 1994; Liu, 2001). Thickness measurement
(Spence and Zuo, 1992), measurement of lattice parameters and strain fields
(Randle et al., 1989), Since CBED uses focused beam the information
retrieved corresponds to a very small area, usually around a few hundred
nanometers. The following gives the important steps to obtain high qual ity on-
zone axis CBED patterns:
( 1) Choose a proper spot size: the smaller the spot size, the less chance
Advanced Transmission Electron Microscopy of Nanostructured. . . 123

100 314
90 410
80
70 214
C 60
00 413
<:: 50
~ 40 311
.5
30
20
10
0

(a)

100
90 2111
80
70
.~ 60
<::
50 2110
]" 40
30
IOIl
t
20
10
0

(b)

Figure 4.9 Electron diffraction patterns of (a) NdFel4 8 2 , (b) Cos Sm. The calculated
intensity is attached to the experimental pattern.

to encounter a defect.
(2) Correct astigmatism of the condenser lens in the image mode by
making the electron beam circular.
(3) AI ign the high voltage center.
124 Y. Liu. J. Sellmyer

(4) In the diffraction mode, align the zone axis by tilting the specimen.
The rough alignment can be judged by watching the relative positions of the
pole of Kikuchi lines to the diffraction pattern. When the pole of Kikuchi lines
and the central beam coincide. the specimen is set at zone axis.
(5) Chose a proper condenser aperture or the ex selector (in the JEOL
TEMs) so that adjacent disks are not overlapping.
(6) The final fine alignment of zone axis can be done either by shifting the
condenser aperture or by tilting the beam. Precise alignment can be done by
watching the central. The thickness contour or other features should be
symmetrical.
(7) An alternative way of zone axis alignment is to translate the specimen
while observing the CBED pattern until an optimum pattern is obtained.
(8) Repeat (6) and (7) until an optimum pattern is obtained.
(9) Expose the film. Always take several exposures with different
exposure time and take both shorter camera length patterns and longer camera
length patterns.
In field-emission TEMs. the spot size has become so small that beam
damage has become significant. It is necessary to use slightly defocused beam
to reduce beam damage. The convergence angle can also be changed
continuously by changing the degree of defocus of the condenser lens.

4.3.3 Nanodiffraction

Nanodiffraction is different from SAD in that it uses a focused beam but is


similar to CBED. The difference between nanodiffraction and CBED is that
CBED requires the specimen be thick enough to generate sufficient contrast in
the diffracted disks while nanodiffraction requires the probe be small enough
for the area being studied. Nanodiffraction has been successfully used to study
the metastable phases in magnetic materials.

4.3.3. 1 Experimental Method

The experimental set up for nanodiffraction is realized by focusing the electron


beam on the specimen as shown in Fig. 4. lOa. The area contributing to
diffraction is controlled by the probe size. When the probe is focused,
minimum probe size can be obtained. However, the diffraction spots in the
diffraction pattern are disks instead of sharp spots. The diameter of the disk is
defined by the semi-convergent angle of the electron beam. This could cause
overlapping of spots if the unit cell is large. The sharp spot pattern can be
obtained by defocusing the probe slightly using the condenser lens as shown in
Fig. 4. lOb. The probe size on the specimen in this set up may be larger than
the available minimum probe size.
Advanced Transmission Electron Microscopy of Nanostructured... 125

(a) (b)

Figure 4. 10 Experimental set-up for nanodiffraction, Ca) focused beam,


Cb) defocused beam.

The experimental procedure to obtain a nanodiffraction pattern is:


(1) Choose a proper probe size: A probe size smaller than the half grain
size under study is preferred. The smaller the probe size, the less chance to
encounter a defect. However, if the probe is too small, the diffraction pattern
would be too dark to be seen. Another point needed to mention is that the
probe size must be a few times larger than the unit cell in order to reserve the
symmetry of the crystal in the diffraction pattern.
(2) Align the high voltage center and current center. Both are important
for obtaining a good probe.
(3) Find the zone axis. Going through the above procedure will generate
a good probe. Now going to the diffraction mode will generate a diffraction
pattern. Although tilting can still be used to find the proper zone axis, large
angle tilting is limited by the size of the grain and can cause overlapping with
adjacent grains. Therefore, finding the zone axis is alternatively done by
probing the right grain by scanning the probe over the specimen while watching
the diffraction pattern, until the one with the desired zone axis is found.
(4) Correct astigmatism of the condenser lens in the image mode by
making the electron beam circular.
(5) Correct astigmatism of the objective lens. This could be done using a
larger spot size. This alignment ensures that a minimum spot size can be
obtained.
(6) Focus the image. This is done by setting the objective current at a
desired value and then adjusting the z control (the vertical translation of the
specimen). The final fine focus can be done by adjusting the objective current.
(7) Expose the film. Shorter exposure time 1- 2 s is preferred. This can
126 Y. Liu, J. Sellmyer

be done by using smaller camera length or using a CCO camera.


Nano-probe placed at crystals can cause very fast contamination or
decrystallization. The contaminants are light elements such as C etc. The
clear diffraction pattern can disappear in a few seconds. The contamination
can come from poor vacuum or the specimen surface. Using a cold specimen
stage is an effective way to reduce specimen contamination. The contaminant
atoms can be "frozen" at the specimen surface. Heating the specimen to
1OOC is also occasionally helpful to eliminate contaminants from the specimen
surface. Another trick to take good quality of nanodiffraction patterns is to
reduce the beam illumination time on the crystal under examination.

4.3.3. 2 Applications of Nanodiffraction

Identification and measurement of volume fraction of amorphous phase:


Fig. 4. 11 is HRTEM image of Co-8m film. The grainy contrast without lattice
fringe in HRTEM image has frequently been considered as an indication of
amorphous phase. This judgement is inaccurate. When a crystallite is aligned
along a high index zone axis, the projection of the structure is beyond the
resolution limit and no lattice fringes will be observed. Contamination of the
crystallite will also result in an amorphous like contrast. The phase
identification of the amorphous phase is most reliably made by nanodiffraction
without the limitation of the microscope resolution. No matter what direction is
aligned relative to the electron beam, a spot pattern is always generated if a
crystal is probed. Figure 4. 12 shows two examples of the nanodiffraction
patterns taken from the Co-8m film. Figure. 4. 12a is a case of the amorphous
phase, and Fig. 4. 12b is the pattern of the crystallite (Liu et aI., 1994). To
measure the volume fraction of the amorphous phase, each fi 1m was probed
with a separation of about 2 nm along a straight line, which was scanned in an
area for 200 measurements. In most cases, either a spot pattern or an

Figure 4. 11 HRTEM image of a Co-8m film.


Advanced Transmission Electron Microscopy of Nanostructured... 127

amorphous pattern was observed. For some cases where a mixture of a spot
pattern and an amorphous pattern is observed, and the pattern is judged as
either a spot pattern or an amorphous pattern depending on the intensity
distribution. The volume fraction of the amorphous phase in the film was
deduced by the ratio of the number of amorphous phase pattern against the
number of total patterns. The results for different films along the coercivity are
listed in Table 4. 1. The volume fraction of the amorphous phase varies with
the processing parameter Ar pressure and the composition. Higher Ar pressure
and higher 8m content result in a high volume fraction of the amorphous phase.

(a) (b)

Figure 4.12 Nanodiffraction patterns of (a) from amorphous phase, (b) from crystal.

Table 4. 1 Film processing, nanostructure and magnetic properties relation in Co-X


(X=Sm,Pr,Pt) films. Vc is the volume fraction of the crystallite. C indicates crystallite and
A amorphous phase.
Film com. Ar Pressure Film thickness Coercivity
Phases Vc
(at. %) (mTorr) (nm) (kOe)
Co-19%Sm 5 24 061 C+A 91%
Co-19%Sm 12 24 258 C+A 65%
Co-19%Sm 30 24 0.92 C+A 54%
Co-22%Sm 5 30 12 C+A 81%
Co-22%Sm 17 30 4.1 C+A 57%
Co-22%Sm 30 30 3.4 C+A 48%
Co-22 % Pr
20 35 0.5 C+A 32%
as-deposited

Co-22 % Pr
20 35 7.2 C 100%
400C 30min

Co-50 % Pt
20 10 0.1 C 100%
as-deposited

Co-50%Pt
20 10 23 C 100%
750C 3h
128 Y. Liu, J. Sellmyer

Crystal structure identification of nano-crystallites: As shown in Fig. 4. 11


the nanostructure of Co-8m film is made of crystallites of about 5 nm distributed
in amorphous matrix. The composition of the film corresponds to Cos 8m,
which has the highest anisotropy value and is an excellent permanent magnet
material. However, anisotropy measurement of the film showed much lower
anisotropy, indicating a different phase. X-ray and electron diffraction studies
indicate that the diffraction data of the sputtered films produced in our
laboratory can not be matched with any equilibrium phases in the Co-8m
system. The crystal structure of the crystallites was identified by
nanodiffraction and confirmed by HRTEM (Liu et al., 1995). It was found that
all the nanodiffraction patterns could be indexed by close-packed structure
models with different stacking modes. Figure 4. 13 compares the experimental
and simulated [1 1 '2 0] zone axis nanodiffraction patterns. In earlier study it
has been shown that [1 1 '2 0] zone axis pattern has the characteristics of the
stacking mode and leads to a decisive identification of the stacking mode (Liu
et al., 1994). The stacking modes ABAB, ABCABC, ABACABAC and
ABCABABCAB are identified. In many cases, a pattern with overlapping spots
was obtained, suggesting a random stacking. The close-packed structure
stacking modes were later confirmed by HRTEM. HRTEM images of two
crystallites are shown in Fig. 4. 14. The simulated HRTEM images are
attached. The stacking mode can be identified directly by the packing of
atoms.

(a) (c) (e) (g)

(b) (d) (f) (h)

Figure 4.13 [1 1 "2 oJ zone axis patterns of Co-Sm films. Upper panel is simulation and
lower panel TEM patterns. (a) and (b) two layer stacking mode ABAB, (c) and (d) three
layer stacking mode ABCABC, (e) and CO four layer stacking mode ABACABAC, (g) and
(h) five layer stacking mode ABCABABCAB.
Advanced Transmission Electron Microscopy of Nanostructured... 129

(a)

(b)

Figure 4.14 [1 1"2 OJzone axis HRTEM images of two crystallites.

4. 4 High Resolution and Super-Resolution TEM

HRTEM has played important role in materials characterization. Viewing


directly the atoms at resolution of 0.17 nm has become reality in commercially
available TEMs. However, because of the defect in magnetic lens in TEM, the
contrast transfer function (CTF) of an objective lens in TEM convolutes the
recorded image in two ways: one is the spherical aberration that changes the
phases of the scattered beams and the other is the dumping envelop that
130 Y. Liu, J. Sellmyer

weakens the amplitudes of beams with high spatial frequencies. The added
phase could change the color of the atoms, resulting in uncertainty of atom ic
positions and cause zero intensity at certain frequencies. Therefore the
resolution of commercially available TEM with accelerating voltage at and
below 400 kV is limited to O. 17 nm. Extending the resolution of TEM to
O. 15 nm or better has been referred to as super-resolution TEM. There are
currently two approaches to achieve super-resolution. One is to use improved
instrument such as scanning transmission electron microscopes, correction of
the objective lens aberration and the use of higher accelerating voltage. The
disadvantages of such approach are that they are very expensive, and limited
to a few laboratories at the present time. High-energy electrons, for example,
can substantially cause radiation damage to the specimen. The second
approach is to use the current generation TEM and image processing to
achieve the same level resolution without high cost or subjecting the specimen
to high-energy electrons. Resolution extension by hollow-cone illumination
(Sidorov et al., 1997; Dings et aI., 1994), direct phase retrieval by off-axis
electron holography (Orchowski et al., 1995), and exit wave reconstruction
using post image processing have been exercised. Among the post image
processing approach, Kirkland et al. developed a process by variation of
objective lens focus (Kirkland and Siegel, 1980; Kirkland, 1982,1984). Hu
et al. combined HRTEM image and electron diffraction pattern to reconstruct
the image (Hu and Li, 1991; Hu et aI., 1992). Side-band method
(Hohenstein, 1991), Maximum-likelihood method (Coene et aI., 1996),
parabola method (Beeck et aI., 1996) were recently developed. Examples of
high resolution images close to O. 1 nm have been reported (Dong et al.,
1992). The common feature of these investigators is that they rely heavily on
post image processing to evaluate the contrast transfer function and use
iteration to achieve the final result. None of the above methods has been made
commercially available yet. In this chapter, we present another route to
extend the resolution of an electron micrograph. The method will help to
understand the principle of image processing. The difference between our
approach and others is that we carefully take micrographs and acquire high
quality signals. Consequently the later image reconstruction is rather
straightforward.

4.4.1 An Image Processing Model

The idea of our image processing is to reconstruct the image as if a perfect


transfer function T p is used. This perfect transfer function T p is defined to be:
CD it transfers all the spatial frequencies with the same additional phase - TT/2
IlX (..1X is allowed error), (2) it does not attenuate the diffracted beams of
any spatial frequencies. CD will be done by retrieving signals with the same
additional phase from a series of micrographs taken under certain conditions.
Advanced Transmission Electron Microscopy of Nanostructured. . . '31

(2)will be done by using the wave amplitude recorded by an electron diffraction


pattern (Hu and Li, 1991; Hu et aI., 1992). The recorded image intensity in
an electron micrograph, under the weak phase object approximation, can be
expressed as
I ( r) = 1 + 2 0-4> ( r) * Fr 1[ T( k , z )] (4.1)

The contrast C(r) is then


C ( r) = 4> ( r) * Fr 1 [T (k, z ) ] (4.2)
where 0- = 1T i\ / V, i\ is the electron wave length, V is the accelerating
voltage, 4> ( r) is the projected potential (PP), r is the position vector,
* represents convolution, FT- 1 represents inverse Fourier transform,
T (k , z) is the contrast transfer function (eTF), k represents spatial
frequency having the dimension (nm- 1 ) , and z is defocus. The eTF is
expressed as
(4.3)
where A (k) defines the aperture function, E s ( k , z) defines the envelope
function due to partial spatial coherence and E t (k , z) defines the envelope
function due to partial temporal coherence. For weak phase object, Tc (k ,z)
is
Tc=sinX(k,z) (4.4)

where X ( k , z) is the phase added by the objective lens. Because of the


modification by T ( k , z), the image has the resolution for structure imaging
named Scherzer resolution. The phases of diffracted beams with spatial
frequency beyond Scherzer resolution are scrambled.
To reconstruct an image in which all the diffracted beams are transmitted
with the same additional phase, a series of images are taken under special
defocus conditions which we determine next. The phase shift by the objective
lens is given by

(4.5)

where C s is the spherical aberration, z is the defocus with under defocus as


positive.
Figure 4. 15 plots X against k. X shows a minimum at k = k m and
changes much slower near k = k m' Signals near k = k m are those that we want
to retrieve. Let d X / dk = 0, we have

(4.6)

Substituting Eq. (4. 6) into Eq. (4.7) yields

(4.7)
132 Y. Liu, J. Sellmyer

To ensure maximum signal transmission, we choose

x=_n-non (4.8)
2

o 8 o 4

-2 -2
~ ~

-4 -4

(a) (b)

~ ...+....------'--+-------t-t-t--------\-;H-f-t'f
.;;; 0
-I

-I -I

(e) (f)

Figure 4. 15 Additional phase X and contrast transfer function sin X as a function of


spatial frequency k. The defocus is 82 nm for (a), (c) and (e) and 109 nm for (b), (d)
and CO. (c) and (d) have no damping. (e) and (f) are calculated using defocus spread 2
nm, semi-convergence angle 1.5 mrad.

where sin X takes maximum amplitude, as the center of signal retrieval. Here
n = 0, 1,2,3, ... When n is even, beams will be transmitted with additional
phase - ; - N 0 2n( N = 0, 1 ,2,3, ... ) and can be treated as having the same
phase. When n is odd, beams will be transmitted with additional phase
- 3n
2 - N 0 2n. In this case n is added after retrieval of such signals. As a
Advanced Transmission Electron Microscopy of Nanostructured... 133

result, signals with additional phases - ; - N 2TT are retrieved for


reconstruction of the image.
Next, we determine the pass - band width Ilk = k 2 - k 1 for signal
collection. In our case, the resolution of the reconstructed image is targeted at
O. 14 nm for the JEOL 2010, and the allowed error of phase shift by the lens is
set under the condition:

I sin (X L\X) I ~ O. 9 (4.9)

with IlX = O. 14TT. Taking - ; - n TT IlX as the width of the signal

collection and substituting X = - ; - n TT - IlX into Eq. (4. 6) to solve for

z, which is denoted as z d' we have

(4.10)

Eq. (4. 10) ensures that the minimum in the X - k curve takes the phase of

X=- ; - n TT-IlX Substituting Eq. (4. 9) into Eq. (4. 10) gives

(4.11)

Substituting X =- ; - n TT + IlX into Eq. (4. 11) and solving for k, one

gets two solutions for k>O which are

_ [- b (b 2 - 40C)05 JO.5
(4.12)
k1.2 - 20
b = - Zd A2 C = ..l (1 + 2 n _ 2 L\X ) (4. 13)
2 4 TT

Signals between k 1 and k 2 have additional phase X =- ; - n TT IlX .

Table 4.2 lists results calculated from Eq. (4.8) to Eq. (4. 12). n represents
different pass-bands, Z d is the defocus at which the micrograph should be
taken, k 2 - k 1 is the width for signal retrieval. There is an overlapping zone
for signal retrieval at adjacent pass - bands. This means that the signal
retrieval could be completed from k =
2. 1 to any high frequency that is
transferred by the objective lens.
In order to retrieve low spatial frequency k<2. 1 nm- 1 , we set the signal
collection zone at k 1 < k < k 2. In this zone the phase should be between

- ; +IlX and - ; -IlX with IlX=0.14TT. From Eq. (4.5) taking defocus
Z as a function of spatial frequency k, one gets
134 Y. Liu, J. Sellmyer

(4.14)

Table 4.2 List of pass bands for image reconstruction. n is the pass band order, The
symbol' indicates that the signal collecting zone is at low frequency. Zd is the defocus. Other
parameters are: Cs =0.5mm, i\=0.00251nm. d,=I/K" d 2 =I/K 2
Zd k, k2 d 2 or d,
n (nm) (nm-' ) (nm-' ) (nm)
,2 117 1.1 15 '0 909
, 1 65 15 2 1 '0 666
0 40 2.1 4.6 0.217
1 64 3.5 5.4 O. 186
2 81 42 5.9 0 171
3 96 4.7 6.2 O. 161
4 108 5.1 65 0 153
5 119 5.4 6.8 0 147
6 129 57 70 0 142
7 138 6.0 7.2 O. 138
8 147 62 7.4 0 135
9 156 6.4 76 O. 132
10 163 6.6 7.8 0 129

Substituting X = - ; - b.X' b.X = O. 14n corresponding to k = 2.1 nm,

i\ =O. 00251 nm and Os = O. 5 mm, one gets z = 65 nm.


Knowing z allows one to solve for k from Eq. (4.5) :

k1,2 -
_ [- b (b 2 - 40C)05
20
r 5
(4. 15)

= Os 4 i\
i\ 2
=~ 2~X)
4 Z
o b =--2-' C (1 - (4.16)

Substituting z = 65 nm, X = - ; + b.x, one gets two roots. k = 1. 5 nm- I

corresponds to k I ' One of the signals collecting zones is then determined to be


1. 5 nm- I <
k<2. 1 nm- I . Similarly one can deduce signal collection zones at
lower spatial frequency by going to higher defocus value. Two zones 1 and
2 for low spatial frequency signal collection are included in Table 4. 2.
To retrieve the amplitudes of the diffracted beams, we first Fourier
transform the intensity of the micrograph In (k ,z) taken at nth pass-band,

FT[ In ( k , z) ] = a + 2 aF (k) T (k, z) (4.17)


Advanced Transmission Electron Microscopy of Nanostructured... 135

Here, 0 is the delta function, 2 of (k) is equivalent to diffracted waves for


crystalline materials and F ( k) is the structure factor F (hkl). The relative
values of F (hkl) can be deduced from electron diffraction pattern by
measuring the intensity I (hkl) of each beam. Then we use F(hkl) =
c [I (hkl) JO 5, where c is a constant to retrieve F ( hkl). Since we are only
interested in the contrast of the reconstructed image, c is taken to be 1 for all
the beams.

4.4.2 Procedure for Image Reconstruction


In summary, the image reconstruction procedure is:
(1) An accurate measurement of C s and defocus step must be
accomplished.
(2) Calculate the defocus setting for each pass-band by Eqs. (4.10) and
(4.14) .
(3) Take a diffraction pattern of the phase.
(4) Take high resolution TEM images at each pass-band.
(5) Fourier transform TEM images and retrieve signals in spatial frequency
zone kz-k l from each image as listed in Table 4.2. The size of each zone is
determined such that the whole range spatial frequency of interest is covered
without overlapping.
(6) Replace the amplitude of each beam with those measured from
electron diffraction pattern.
(7) Reconstruct the image by inverse Fourier transform of the retrieved
signal zone.

4.4.3 Test of the Image-Processing Model


To understand the image processing a detailed reconstruction is shown in
Fig. 4. 16 for the case of [110J zone axis of GaAs. The first row shows
images decomposed from a single TEM micrograph. The second row shows
images with the right phase and amplitude. The final image is the sum of the
decomposed images. Both the amplitude and the phase affect the final image.
A concern is whether such a simple reconstruction procedure would work
for defect structure or complicated structure at higher resolution. Simulation
was performed to explore the validity of the method. [110J zone axis of a
rather large unit cell of an important magnetic phase Fe14 Ndz B with 68 atoms
was chosen. Figure 4. 17a, b, c compares projected electrostatic potential,
reconstructed image by the current method and simulated image assuming a
TEM with resolution of O. 069 nm. The pass-bands used for image
reconstruction are given in Table 4. 2. In both cases, the image is a projection
of the structure. This shows that indeed the reconstruction method proposed
136 Y. Liu, J. Sellmyer

Sum {Ill} {202} {113} {004}

Figure 4.16 Illustration of image reconstruction of GaAs in [110J direction. First row
shows images decomposed from a single TEM micrograph. The second row shows images
with the right phase and amplitude. The image on left most of each row is the sum of the
decomposed images.

can be applied to any complicated structure or defect structure (defect


structure is nothing but a structure with a much larger unit cell). The sl ight
difference in contrast between the two images is due to the transfer functions.
The transfer function for Fig. 4. 17c attenuates low spatial frequencies more
than the transfer function for Fig. 4. 17b. Therefore, Fig. 4. 17c shows lower
average intensity than Fig. 4. 17b.
In this approach we have made use of the weak phase approximation.
This requires the specimen to be very thin. The successful image
reconstruction for GaAs sample indicates such thin samples are readily
obtainable using the standard ion milling technique. A second issue is whether
a single image or focal series is preferred. For metallic materials, our
experience shows that the sample could at least remain stable for 30 min for
HREM observation. Therefore, taking more images by the use of CCO camera
is not a problem. Rather the maximum transmission of signals is more
important.
In summary we have proposed a new method to reconstruct TEM image
with extended resolution. Because of its simplicity, resolution at the O. 15 nm
level is routinely achieved using the JEOL 2010 TEM with La6 B filament. The
features of the current method are summarized in the following.
( 1) The use of electron diffraction for retrieval of amplitude is more
straightforward than modeling the damping envelope and therefore is adopted
in our model.
(2) The use of a single image for reconstruction, which could be
important for biological samples, has large errors near zeros of the transfer
function. We therefore employed the focal series method.
(3) In all the earlier focal series models, the signals transmitted by pass-
bands where the transmission of ampl itude is close to 100 %, and the signals
Advanced Transmission Electron Microscopy of Nanostructured... 137

.. .
. ..
~. ..... .. ... ...'

...
.. . ..

."
'. . ~ .

...
' '~

.,.
.-. ..
- .. . ,.
" ,

.. .. .. ...
-. ... .-.... ~.

(a)

(b)

(c)
Figure 4. 17 Comparison of [110] zone axis images of Fel4 Nd2 B. (a) projected
electrostatic potential, (b) reconstructed image by the proposed model. The spherical
aberration C s is O. 5nm and the final image is reconstructed from 15 images at defocus
setting listed in Table 4. 2. (c) simulated image assuming a TEM with C s =0. 01 mm and
Scherzer resolution of O. 06 nm. For both simulations (b) and (c) the other parameters
are: A = O. 00251 nm, semi-convergence angle = O. 01 mrad, defocus spread = O. 01 nm,
specimen thickness = 1. 6 nm. The cut off spatial frequency is 14.6 nm- 1 corresponding to
resolution of O. 069 nm.
138 Y. Liu, J. Sellmyer

transmitted outside the pass-band where the amplitudes are seriously damped
and the phases change drastically with spatial frequency, are treated equally
and therefore reduced the signal-to-noise ratio. We, on the other hand, simply
cut off all the signals outside the pass-band. The result is that the signal-to-
noise ratio is significantly improved and that the reconstruction of image
becomes straightforward.

4.5 Selected Reflection Imaging

4. S. 1 Origination of the Technique

In the research on high-density recording media based on hexagonal close-


packed Co alloy films, one wishes to know how many grains have the c-axis in
the film plane, which give rise to the in-plane anisotropy. So far there has
been no direct way to measure this quantity. X-ray diffraction has been used to
determine the texture of bulk materials. However for thin films in the order of
10 to 30 nm, X-ray diffraction peaks can be very weak or not observable. In
electron diffraction pattern, it is noted that when the c-axis of a grain lies in
film plane, it produces a diffraction spot (0002). As a result, all the grains
having the c-axis in film plane (0002) parallel to electron beam contribute to
the diffraction ring (0002). If this (0002) ring is used to form the image, only
those grains with the c-axis in film plane will have high intensity. Since the
selected ring corresponds to a particular reflection, it is therefore named
selected-reflection imaging (SRI). Table 4. 3 compares the image contrasts
for different modes. One outstanding feature of SRI is that the intensity of a
grain is a strong function of the crystallographic orientation relative to the
electron beam. This feature gives SRI many unique applications.

Table 4.3 Comparison of selected reflection image, bright field image and annular dark field
image contrast: where F is the structure factor of the reflection, t is the thickness of the
particle, s is deviation parameter from Bragg reflection, a function of the crystal orientation
relative to the electron beam. 8 0 is aperture angle, Z is atomic number, Rand aH are
related to the collection range of signal and characteristic scattering, 5 is the ratio of
effective scattering cross section to total cross section.
Selected reflection imaging Bright field Annular dark field
2
1,= (F si:~tsf over all the rings I
a
=-.l82z2
3 0 R
a~
"
V
I b =l- ~ I,
High contrast, high signal, High resolution
most sensitive to crystal Sensitive to thickness, Sensitive to thickness and
orientation and orientation atomic species
Advanced Transmission Electron Microscopy of Nanostructured... 139

4.5.2 Experimental Method in Conventional TEM


The most simple set up for obtaining a selected reflection imaging is to place
an annular aperture in the back focal plane of the objective lens to allow the
selected ring to go though as shown in Fig. 4. 18a. In real operation, instead
of using a annular aperture, use is made of the tilting of the illumination as
shown in Fig. 4. 18b. A small aperture is used to allow a portion of the ring
selected to go through the aperture to form a dark field image. The film is
exposed for a proper time under this condition. The illumination is then tilted
such that the next portion of the ring goes though the aperture to continue the
exposure. This process is continued until the whole ring is covered. The final
image is the superposition of dark field images formed by all the beams in the
whole ring. The tilting of the beam can be easily done using the following
program for the JEOL 2010.
Selected reflection ring

(a)

Selected reflection ring

Aperture
2 3

(b)

Figure 4. 18 Experimental set-up for SRI. (a) an annular aperture is used to select the
ring. (b) beam tilt is used to select the ring by exposing the ring part by part.
140 Y. Liu, J. Sellmyer

4.5.3 Application of SRI

(1) Measurement of grain size of nano-particles (Liu et ai., 1998): The grain
sizes of nanostructured materials are usually measured by bright field TEM
image, high resolution TEM image or annular dark field image. In the above
three modes, the image is a projection of the grains in the specimen. This is
not a problem if the specimen is very thin or the particles are distributed on the
supporting film without overlapping. However, for nanoparticles in the range of
1 to 5 nm range, the particles tend to agglomerate. This limits the
effectiveness of all the conventional techniques. Figure 4. 19 compares the
bright field image and a selected reflection image taken at the exact same
region from Co particles synthesized by chemical method. The contrast in
Fig. 4. 19 is significantly enhanced. Another unique property of SRI is the
ability to image the small particles even they are overlapped with another big
particle. Fig. 4.20 is an example. A large particle in Fig. 4. 20a blocks all
other small particles that are overlapping with it, while by using the selected
reflection imaging in Fig. 4. 20b, small particles that are overlapping with the
large one can be revealed.

(a) (b)

Figure 4.19 Comparison of (a) bright field image and (b) SRI showing the small Co
particles.

(2) Measurement of volume fraction of second phase in-dual phase


material: Sometimes a bright TEM image or high resolution TEM image does
not reveal the difference between two phases. The ex istence of two phases
can be confirmed by the electron diffraction pattern. By using a ring
corresponding to a particular phase, that phase could be identified. Two
phases, an ordered L 10 phase and a disordered fcc phase were identified by
Advanced Transmission Electron Microscopy of Nanostructured... 141

(a) (b)

Figure 4. 20 Comparison of (a) bright field image and (b) SRI showing the small
particles of Pt particles.

(a)

(b) (c)

Figure 4.21 SRI image: (a) (002) reflection image and (b) (001) reflection image of
FePt film (c).
142 Y. Liu. J. Sellmyer

using (001) image and (002) image (Liu et al., 2000). This example is shown
in Fig. 4. 21. Figure 4.21a is the (002) image showing the grains with the
preferred direction and Fig. 4. 21 b is (001) reflection image. The bright grains
in Fig. 4. 21 b correspond to the ordered L lO phase while those that are white in
Fig. 21a but dark in Fig. 21b are Fe solid solution. This method is most
sensitive to the ordering disregarding the chemical composition of the phases.
(3) Texture identification: in longitudinal magnetic recording media, it is
very important to have the c-axis aligned in the film plane. The grains with c-
axis in film plane can be found by a selected reflection (0002) image.
Fig. 4. 22a shows the diffraction pattern of a' metastable phase C0 3 Sm. This
phase is an ordered hexagonal close-packed (hcp) structure. However,
because the (2"20) ring, (0002) ring and (2121) are too close to each other,
it requires a high angular resolution and the use of a very small aperture (0. 5
IJm), which is not practical. Very small aperture also reduces the resolution in
the image. The solution to this limitation is given in the next section.
100 2021
S 90
S 80
B 70
"'-;"
<l)
60
u
~ 50
g 40
'"<:: 30
E 20
.E
10
0

Figure 4. 22 Diffraction pattern of C03 8m.

4. S. 4 Experimental Set-Up in a STEM


Two parameters are used to characterize the performance of SRI: image
resolution and angular resolution. In a conventional TEM, the image resolution
d using a small objective aperture is defined by diffraction limit:
Advanced Transmission Electron Microscopy of Nanostructured... 143

d = J..L (4.18)
2A r
where J.. is the wave length of the electrons, L is the focal length of the
objective, A r is the radius of the aperture. The angular resolution is defined by
the minimum angle between two adjacent rings that can be effectively
selected. The angular resolution e is given by

e = C zAL-r (4.19)

where Cz is a constant close to 1. It is clearly seen that higher angular


resolution is achieved by the expense of image resolution. For 200 kV
electrons, for example, JEOL 2010 (A = O. 00251 nm) and L = 1 mm, to select
the (0002) ring in hcp Co, angular resolution of 2.89 mrad is required. This
corresponds to an aperture radius of 1. 44 IJm and image resolution of 1. 74 nm.
This resolution is satisfactory for most cases. However, for small phases with
large unit cell, such as the magnetic phase Ndz Fe14 B shown in Fig. 4. 9a, high
angular resolution of the order of 1. 5 mrad is required. This would require an
aperture of radius O. 8IJm resulting in image resolution of 3.48 nm. Such performance
is not acceptable for imaging a material with a grain size of 5 - 10 nm.
A convenient solution to this problem is to use scanning TEM (STEM). A
thin annular dark field detector is used to select the reflection ring. This could
be conveniently done by placing an annular aperture above a conventional
annular dark field detector. The aperture is painted with fluorescence for the
observation of the diffraction ring. The camera length can be adjusted by the
free lens control unit. The alignment of the ring (X-Y shift) can be adjusted by
the deflector just as in the diffraction mode.
In STEM, the image resolution is mainly determined by the probe size.
The current FEG TEM can generate 0.2 nm probe. The angular resolution is
determined by the convergence of the electrons. Standard FEG TEM can
generate a probe with semi-coherence angle of about 1 mrad. Angular
resolution can be further enhanced by using a slightly defocused beam with
minor reduction in resolution. Since selected reflection imaging retrieves the
reflection signal which is generated according to the Bragg's law, the probe
size should be at least as large as two atomic columns. The probe size and
therefore the resolution of selected reflection imaging using a STEM is limited
to O. 4 nm. The angular resolution can be as high as 1 mrad.

4. 6 STEM and Z -Contrast Imaging

Principles of STEM: in contrast to CTEM which is a parallel process for the


image formation, STEM is a serial process of the image formation. The
144 Y. Liu, J. Sellmyer

electron beam is focused into a small probe by the objective lens, which is
located before the specimen. When acquiring the image the probe is scanned
over the specimen and the transmitted electrons are detected by various
detectors corresponding to various angular ranges. The intensity at the image
(x' , y') is generated by the probe at the specimen position (x, y). The
magnification is gained by di/ds where di is the distance at image between
(x; , y;) and (x;, y;) and ds is the distance between (XI' Y1) and (xz, yz)
at the specimen. Various detectors can be used to collect the signal: Bright
field detector has a small axial collector aperture which detects low angle
electrons. If minor inelastic scattering is ignored, the bright field STEM
imaging optical set-up is equivalent to conventional TEM according to the
principle of reciprocity (Cowley, 1969). Annular dark field detector (ADFD)
is a donats shape detector and was introduced by Crewe, et al. (1970). The
first success of ADFD was the image of single heavy atoms by Isaacson et al.
( 1979). Pennycook and Boatner (1988) demonstrated that the STEM image of
thick crystalline materials in a zone axis acquired by ADFD can be explained in
terms of an incoherent image. The strong dynamical diffraction of the crystal
was explained by Pennycook and Jesson (1990).
Z -contrast is a name given to the image formation mechanism in a
scanning transmission electron microscope. This technique has shown promise
in improving the resolution of TEM and chemical analysis at the highest spatial
resolution.
Advantages of Z -contrast imaging: Firstly, the Z -contrast imaging
provides directly interpretable images at atomic resolution. The intensity is
given by the convolution of the electron probe intensity profile with a real and
positive specimen object function. Such CTF never reverses the contrast from
the true image for any spatial frequency. Secondly the majority signals are the
Rutherford scattering from the nuclei peaked at the atomic positions and
proportional to the square of the atomic number. Because of this, incoherent
imaging by ADFD is referred as Z -contrast imaging. This gives Z -contrast
imaging the unique ability to directly observe segregation of impurities,
chemical order and disorder, and atomic column occupancy. Thirdly, the Z-
contrast imaging shows very little change in contrast with increase in specimen
thickness while in coherent imaging dynamical diffraction with increasing
specimen thickness causes contrast reversals.
Composition mapping by EDX in STEM: EDX or EDS stands for energy
dispersive X-ray spectroscopy which has been a standard technique to probe
the composition of small phases in TEM. In STEM the X-ray intensity at each
position can be used to form an image to represent the content of each
element. Such method is useful to understand the segregation of certain
element. The strength of EDX for compositional map is its ability to examine
thicker samples at lower spatial resolution compared to EELS mapping which
offer much higher spatial resolution but require thinner sample.
Advanced Transmission Electron Microscopy of Nanostructured... 145

4. 7 Electron Energy Loss Spectroscopy

When electrons pass through the specimen some are elastically scattered by
the nuclei, reserving their energy and some lose their energy by interacting
with other electrons. in the specimen and become inelastically scattered
electrons. The events for inelastically scattered electrons include phonon 0-
O. 02 eV, plasmons 5 - 25 eV, inner shell loss edge 13 - 2 keV, inter/intra
band transitions 5 - 25 keV. Inelastically scattered electrons cause achromatic
aberration and blur the image in a conventional TEM image. However,
inelastically scattered electrons are also signatures of the atoms present in the
specimen. Using the inelastically scattered electrons to retrieve the
information in the specimen has come to the recognition of electron energy loss
spectroscopy (EELS). EELS has become an important technique for
nanostructure characterization. Electron energy loss spectrometers are
designed to perform EELS. Following summaries three important applications
of EELS.
Zero loss imaging: As the inelastically scattered electrons are the source
of achromatic aberration, a simple application of EELS is to remove all the
inelastically scattered electrons by using the zero-loss electrons to form the
image. Such technique can be used for both image mode and diffraction mode.
The enhanced contrast has enhanced the resolution in HRTEM and CBED
quality.
Thickness map: In order to perform energy filtering, the electrons used
for analysis must only scattered once to carry the information characteristic of
the electronic structure in the specimen. This requires the thickness of the
specimen t < i\ where i\ is the mean-free path of the electrons in the
specimen. t/ i\ can be estimated by
(4.20)

where It is the total intensity and 10 the zero loss intensity. Equation (4. 20)
can be made for every pixel of the image and then the image is equivalent to a
thickness map. When t/ i\ is smaller than 1, elastically scattered electrons are
dominating and multi-scattering is insignificant. This is an important factor for
meaningful elemental map construction by EELS.
Compositional map: As the energy loss of electrons are directly
associated to the inner shell structure, using a energy slit corresponding to a
particular energy loss to form the image will provide the concentration of the
element. The three window method (Jeanguillaume, 1978) has been
incorporated to the digital micrograph, a matured software by Gatan
Corporation for energy filtering. The first two windows are used to estimate
146 Y. Liu, J. Sellmyer

the background. The third window is used for the selected energy filtering,
providing the compositional information. Figure 4. 23 shows the nanostructure
observed in high temperature permanent magnet Sm2 (CoFeCuZr) 17 Three
phases: the matrix, the grain boundary phase and Zr rich precipitae are
identified. HRTEM shows the Precipitate has larger unit cell and is coherent
with the matrix phase. For easy alignment of the zone axis the specimen was
aligned under magnetic filed and then sintered. The specimen has the TEM foil
normal cut parallel to the c-axis (magnetic soft axis). Since the sizes of both
grain boundary phase and the precipitate are a few nanometers EELS mapping
is the only technique offering sufficient spatial resolution. To perform a reliable
image filtering, it is useful to collect the spectrum and examine the edges for
the elements to be analyzed. Figure 4. 24 is an example of the whole spectrum
and an enlarged part showing the edges of interest. After confirming the
position of the edges, image filtering can be performed at the optimum energy
corresponding to the edges. Figure 4. 25 is the compositional map of the same
high temperature permanent magnet Sm2 (CoFeCuZr) 17. The elemental
distribution in the three phases is clearly revealed.

120
8
g 100

~
x 80
en
E
;::l 60
0
u
Ci 40
U
u 20

0
200 400 600 800 1000
Energy loss (eV)
(a)

708 off set.piet


35
-Co-L2,3
80 30 I
S
~
x
~

0
u
Ci
U
u
5

0
800 1000 1200 1400 1600
Energy loss (eV)
(b)

Figure 4. 23 Bright field and HRTEM image of high temperature permanent magnet
8m2 (CoFeCuZr) 17.
Advanced Transmission Electron Microscopy of Nanostructured... 147

/
Snm

Figure 4.24 EELS Spectrum of Sm2 (CoFeCuZr) 17.

(a) (b)

(e) (d)

(e) (f)

Figure 4. 25 EELS compositional map of Sm2 (CoFeCuZr) 17. (a) bright


field image, (b) Co, (c) Fe, (d) Cu, (e) Sm, (I) Zr.
148 Y. Liu, J. Sellmyer

4.8 Concluding Remarks

Accompanying the progress of magnetic materials research, structural


characterization techniques become increasingly important. In particular,
major recent progress in magnetic materials research has been nanostructured
magnetic materials promoted by the demand of ever increasing recording
density of computer hard disks. Transmission electron microscopy is the single
powerful technique that can fulfill all the tasks demanded by a materials
scientist. This includes the determination of crystal structure, defect structure,
grain boundaries, compositional fluctuation and magnetic domain structure.
Many techniques have matured and others are in progress to meet the new
challenges. Among the challenges we address the following issues that will
motivate further focused effort.
Specimen preparation: TEM specimen preparation remains a tedious task
for all microscopists especially for magnetic samples which are usually more
brittle. In general the information offered by TEM has been faithful
representation of the bulk materials. However as the resolution is pushed
towards sub-Angstrom the thickness of the specimen is also required to be
thinner, which will soon reach its limit. The atoms at the surfaces are
bombarded by ions in the case of ion milling and are displaced from equilibrium
positions as compared to those inside the specimen. Such a damaged layer
has little effect to the exit wave function of electrons when the specimen is
thick but will modify it as the damaged layer occupy a large portion for very
thin specimens.
Magnetic field effect of the objective lens: The magnetic field near the
objective lens can reach 2 T or higher. This will destroy the domain structure of
most magnetic specimens. The objective lens is switched off for domain
observation. Doing this usually reduce the resolution of the TEM significantly.
A Os corrector allowing large space at the specimen area and near zero
magnetic filed will have significant impact on improving the performance of
TEM for magnetic nanostructure characterization.

References
Beeck, M. 0., D. V. Dyck and W. Coene. Ultramicroscopy 64: 167 - 183
(1996)
Broeder, F. J. A., D. Kuiper, A. P. van de Mosselaer and W. Hoving. Phys.
Rev. Lett. 60: 2769 (1988)
Broeder, F. J. A., W. Hoving and P. J. H. Bloemen. J. Magn. Magn. Mat.
93: 562 (1991)
Advanced Transmission Electron Microscopy of Nanostructured. . . 149

Bruno, P. J. Appl. Phys. 64: 3153 (1988)


Buxton, B. F., J. A. Eades, J. W. Steeds, G. M. Rackham. Philos. Trans.
R.Soc. (London) 281: 171(1976)
Carcia, P. F. J. Appl. Phys. 63: 5066(1988)
Chappert, C. and P. Bruno. J. Appl. Phys. 64: 5736 ( 1988)
Coene, W. M. J., A. Thust, M. O. Beeck and D. V. Dyck. Ultramicroscopy
64: 109 - 135 (1996)
Cowley, J. M. Appl. Phys. Lett. 15: 58 (1969)
Dinges, C., H. Kohl, H. Rose. Ultramicroscopy 91-100 (1994)
Dong, W., T. Baird, J. R. Fryer, C. J. Gilmore, D. D. MacNicol, G.
Bricogne, D. J. Smith, M. A. O'Keefe and S. Hovmoller. Nature 355:
605 - 609 (1992)
Hohenstein, M. Ultramicroscopy 35: 119 - 129 (1991)
Hu, J.J. andF.H. Li. Ultramicroscopy 35: 339-350 (1991)
Hu, J. J., F. H. Li and H. F. Han. Ultramicroscopy 41: 387 - 397 (1992)
Isaacson, M. S., M. Ohtusuki and M. Ultaut. In: J J. Hren, J. I. Goldstein,
and D. C. Joy. eds. Introduction to analytical Electron Microscopy New
York, Plenum Press, 343 (1979)
Kirkland, E.J. and B. M. Siegel. Ultramicroscopy 5: 479-503 (1980)
Kirkland, E.J. Ultramicroscopy 15: 151-172 (1984)
Kirkland, E.J.Ultramicroscopy 9: 45-64 (1982)
Liu Y. and J. Mazumder. Proceedings of TMS Symposium: Laser Materials
Processing, 69 - 90 (1994)
Liu, Y., A. Eades and J. Mazumder. Ultramicroscopy 56: 253 (1994a)
Liu, Y., J. Mazumder, and K. Shibata. Metall. Mater. Trans. A 25: 37 - 46
(1994b)
Liu, Y., J. Mazumder and K. Shibata. Metall. Mater. Trans. A 25: 487
(1994c)
Liu, Y., B. W. Robertson, Z. S. Shan, S. Malhotra, M. J. Yu, S. K.
Renukunta, S. H. Liou and D. J. Sellmyer. IEEE Trans. Magn. 6: 4035-
4037 (1994d)
Liu, Y., D.J. Sellmyer, B. W. Robertson, Z. S. Shan and S. H. Liou. IEEE
Trans. Magn. 31: 2740-2742 (1995)
Liu, Y., Z. S. Shan and D. J. Sellmyer. IEEE Trans. on Magn. 32: 3614-
3616(1996)
Liu, Y., Z. S. Shan and D. J. Sellmyer. J. Appl. Phys. 81: 5061 - 5063
( 1997)
Liu, Y. Proc. of the 14th International Congress on Electron Microscopy.
Cancun, Mexico, 1, 629 (1998)
Liu, Y. and D. J. Sellmyer. Proc. of Microscopy and Microanalysis. Atlanta,
Georgia, 752 (1998)
Liu, Y., R. A. Thomas, S. S. Malhotra, Z. S. Shan, S. H. Liou and D. J.
Sellmyer. J. Appl. Phys. 83: 6244 - 6246 (1998)
Liu, Y. X. Zhang and Z. Zhang, eds. In: Progress in transmission electron
150 Y. Liu. J. Sellmyer

microscopy. Tsinghua University Press and Springer, 273 (2001)


Liu, Y., M. Zhen, H. Zeng, D. J. Sellmyer. Handbook of nanophase and
nanostructured materials. Tsinghua University Press and Kluwer Academic
Publishers(2002)
Mansfield, J. J. Electron Micro. Tech. 13: 3 (1989)
Orchowski, A . W. D. Rau and H. Lichte. Phys. Rev. Lett. 74: 399 - 402
( 1995)
Pennycook, S. J. and L. A. Boatner. Nature 336: 565 (1988)
Pennycook, S. J. and D. E. Jesson. Phys, Rev. Lett. 64: 938 (1990)
Randle, V . I. Barker and B. J. Ralph. Electron Microsc. 13: 51 (1989)
Roberts, S, J. McCaffrey, L. Giannuzzi, F. Stevie and N. Zaluzec. In: X.
Zhang and Z. Zhang, eds. Progress in transmission electron microscopy.
Tsinghua University Press and Springer, 301: (2002)
Scherzer, O. J. Appl. Phys. 20: 20 (1949)
Sidorov, M. V., M. R. McCartney and D. J. Smith. Proceedings of Microscpy
and Microanalysis. Cleveland, Ohio, 1191- 1192 (1997)
Spence, J. C. H. and J. M. Zuo. Electron Microdiffraction. Plenum Press,
New York, 169 (1992a)
Spence. J. C. H. and J. M. Zuo. Electron Microdiffraction. Plenum Press,
New York. 85 (1992b)
Stadelmann , P. A. Ultramicroscopy. 21: 131 - 146 (1987)
Tanji, T., M. Maeda, N. Ishigure, N. Aoyama. K. Yamamoto, T. Hirayama.
Physical Review Lett. 1038 - 1041 (1999)
Vokel, E. Proc. Microscopy and Microanalysis. Quebec City, Canada (2002)
Zeng, H., M. Zheng, R. Skomski, Y. Liu and D. J. Sellmyer. J. Appl.
Physics 87: 4718-4720 (2000)

We are grateful to our colleagues Z. S. Shan, C. Chen. C. Nelson. U. Dahmen. S. H. Liou.


J. P. Liu, and R. Skomski for helpful discussions. We are indebted for financial support to
NSF-MRSEC <DMR-0213808). NSF-DMR 0076504, DOE. ONR. AFOSR. ARO. NRI and
CMRA.
5 Mossbauer Spectroscopy Characterization of Soft
Magnetic Nanocrystalline Alloys

Michat Kopcewicz

5. 1 Introduction

With the continuing demand for increasingly efficient magnetic materials for
technological applications, the theoretical and experimental effort aimed at the
development of novel magnetic materials has increased dramatically in recent
years. Of particular interest are artificial materials whose properties are
governed by reduced dimensionality in the nanometer length scale, e. g., soft
magnetic nanocrystalline alloys, metallic multi layers exhibiting the giant
magnetoresistance effect, nanopowders, and rapidly quenched permanent
n1agnets.
Nanocrystalline materials are single-phase or multiphase polycrystals,
with crystal size in the nanometer range, typically 5 to 50 nm. Such materials
can be produced by various methods, e. g., compaction of nanometer-size
powders, deposition techniques or by crystallization of amorphous precursors.
Due to very small dimensions, nanocrystalline materials contain large volume
fraction of grain and interface boundaries which may significantly alter
physical, chemical and mechanical properties as compared to the conventional
coarse-grained polycrystalline materials. Nanocrystalline materials may reveal
improved hardness and ductility, reduced elastic modulus, enhanced
diffusivity, enhanced thermal expansion and superior soft or hard magnetic
properties. Therefore they provide us an excellent opportunity to study the
artificial magnetic materials whose properties are directly related to the
refinement of the grains in the nanometer length scale.
In this chapter our interest will be focused on a new class of soft magnetic
materials, important for technological applications and attractive from the
point of view of basic research, that has been developed in late 1980s. It has
been reported that the bee structure with a nanoscale grain size can be formed
in amorphous FeSiB-based alloys containing Cu and Nb by utilizing the first
stage of the crystallization process. The first experiments performed for the
Fe735Cu\Nb3SiI35Bg alloy (Yoshizawa et al., 1988) have shown that annealing
the amorphous alloy at temperatures between 520'C and 650 'c leads to the
formation of the nanocrystalline bee-Fe (Si) phase with well defined
152 Michat Kopcewicz

homogeneous grains, with a typical size of 10 to 15 nm, embedded in the


retained amorphous structure. The formation of the bcc-Fe (Si) nanostructure
in the FeCuNbSiB system was explained by a combination of an increased
nucleation rate of the bcc phase resulting from the immiscibility of Cu to Fe and
a reduced crystal growth rate due to the small diffusivity of Nb in Fe. This
nanocrystalline phase reveals excellent soft magnetic properties: low coercive
field, high saturation magnetization, high permeabil ity and vanishing
magnetostriction. However, the saturation magnetization is lower than that
typical for the Fe-metalloid amorphous alloys, mainly because of the lower Fe
content related to the addition of Nb and Cu. Excellent soft magnetic
properties are attributed to the decrease of the effective magnetic anisotropy
which is randomly averaged by the exchange interaction and is related to the
refinement of the grain size (Herzer, 1990, 1991). Since the first experiments
by Yoshizawa et al. (1988), considerable experimental effort has been
devoted to the study of structural and magnetic properties of FeCuNbSiB alloys
(Yoshizawa et aI., 1988, 1994; Sui et aI., 1994; Skorvanek and Gerling,
1992; AIIia et aI., 1993; Kulik et al., 1994).
A similar phenomenon was observed also in simpler ternary Fe-M-B alloys
(M: Zr, Nb, Hf, Ta) (Suzuki et aI., 1991a, 1991b, 1993, 1994; Kim et aI.,
1994) which reveal magnetic properties superior to those of FeCuNbSiB alloys
(higher saturation magnetization and permeability). Annealing of the
amorphous precursor causes the formation of nanoscale grains of the bcc-Fe
which exhibit high saturation magnetization (1.7 T). Addition of 1% - 2 % Cu
causes, as before, a decrease of the crystall ization temperature and an
increase of the nucleation rate. However, the Fe-rich Fe-Zr and Fe-Hf alloys
exhibit an Invar effect accompanied by the decrease of the Curie temperature
which leads to the decrease of magnetization at room temperature. A
nanostructure consisting of metastable bcc grains formed in the FeZrB or FeHfB
systems by crystallization of the amorphous phase is expected to exhibit high
saturation magnetization (Makino et aI., 1997a, 1997b). Furthermore, the
solid solutions such as Fe-Zr and Fe-Hf supersaturated phases reveal very
small magnetostriction (-1 x 10- 6 ) (Kataoka et al., 1989).
Recently a new FeCoZrBCu alloy 'with high saturation magnetization and
permeability at temperatures exceeding 600 'c has been developed by Willard
et al. (1998, 1999). In such alloys the nanocrystalline ex and ex'-FeCo phases
are found with significantly improved high temperature properties as compared
with the above mentioned alloys.
Structural properties of nanocrystalline alloys were studied by various
methods including differential scanning calorimetry (DSC) and differential
thermal analysis (DTA) (crystallization temperature), X-ray diffraction (XRD)
and Mbssbauer spectroscopy (for phase identification) and transmission
electron microscopy (TEM) (for determination of the grain size), while the
magnetic properties were usually determined by such conventional instruments
as the vibrating sample magnetometer (VSM), a DC B-H loop tracer and a
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 153

vector impedance analyzer.


Nanocrystalline alloys exhibiting excellent soft magnetic properties are
promising candidates for wide applications in technology. Not only do the
saturation magnetization and permeability exceed those of most conventional
materials (e. g., ferrites, Si-steels) but remanence is high, coercivity is very
small, magnetostriction is almost vanishing and core losses in a wide
frequency range are much smaller as compared with Si-steels or even
Fe-based amorphous alloys (Makino et aI., 1997a). Such a combination of
magnetic properties is very favorable for technical applications in power
transformers, data communication components, pulsed transformers, choke
coils, magnetic heads, sensors and magnetic shielding.
M6ssbauer spectroscopy was successfully used for identification of the
phases formed due to annealing of the amorphous precursors of both the
FeCuNbSiB (Hampel et aI., 1992; Pundt et aI., 1992; Zhou et aI., 1993;
Cserei et al., 1994; Miglierini, 1994; Pradell et al., 1995) and various Fe-M-
B( Cu) - type alloys (Duhaj et aI., 1996; Kopcewicz et aI., 1995a, 1996,
1997a, 1997b; Kim et al., 1996; Miglierini and Greneche, 1997a, 1997b).
However, information regarding the grain size cannot be obtained from the
conventional M6ssbauer measurements and must be obtained either by the
TEM or XRD method. Another method that can provide microstructural
information is small-angle X-ray scattering (SAXS) and this technique has
been applied recently to the nanocrystalline alloys of interest here (Kopcewicz
et aI., 1996, 1997a).
The crystallization behavior of amorphous FeCuNbSiB alloys was studied
in great detail by many researchers (e. g., Hampel et aI., 1992; Pundt et aI.,
1992; Zhou et aI., 1993; Cserei et aI., 1994; Migl ierini, 1994). The analysis
of the M6ssbauer spectra allowed the identification of the nanocrystalline
phase as the bcc-Fe(Sj). Crystallization kinetics was studied for FeCuNbSiB
alloy (Graf et aI., 1996).
New ternary FeMB alloys also attracted a lot of attention. In particular the
structural properties of the nanocrystalline FeZrB(Cu) system was investigated
(Duhaj et aI., 1996; Kopcewicz et aI., 1995b, 1995c, 1996, 1997a, 1997b;
Kim et aI., 1996; Miglierini and Greneche, 1997a, 1997b). Great effort was
devoted to the interpretation of the properties of the interface regions,
separating bcc grains from the amorphous matrix. Miglierini and Greneche
(1997a, 1997b) performed a systematic study of this problem.
Conventional M6ssbauer spectroscopy allows identification and estimation
of the relative abundance of phases formed due to annealing of the amorphous
precursor but do not provide information regarding the magnetic anisotropy,
magnetostriction and size of the grains. Recently the unconventional
M6ssbauer technique which makes use of the effects induced by an external
magnetic radio-frequency (rf) field applied to the material investigated was
154 Michal Kopcewicz

used to elucidate the properties of nanocrystalline FeCuNbSiB (Kopcewicz


et al., 1994a), Fe-Zr-BCCu) alloys (Kopcewicz et aI., 1995b, 1995c, 1996a,
1997a, 1997b) and Fe-M-B (Cu) (M; Nb, Ti, Ta, Mo, ... ) alloys
(Kopcewicz et aI., 1999a, 1999c, 2002; Grabias et aI., 2000; Miglierini
et al., 1999) . The rf-M6ssbauer technique is very sensitive to small changes
of the magnetic anisotropy fields, thus allowing us to distinguish the
magnetically soft nanocrystalline phase from the microcrystalline grains formed
in the course of thermal treatment of amorphous precursors. The rf collapse of
the magnetic hyperfine structure permits the separation of the magnetic dipole
and electric quadrupole hyperfine interactions in the ferromagnetic state by
averaging to zero the magnetic hyperfine field. Using this technique it is
possible to study directly the dependence of the short range order on, e. g.,
the composition of amorphous alloys via the electric quadrupole hyperfine
interaction which is determined by the local environment of the M6ssbauer
nuclei. The rf-M6ssbauer technique was used to study both soft magnetic
properties in various nanocrystalline alloys (Kopcewicz et al., 1996, 1997a,
1997b) and the short range order in amorphous alloys (Kopcewicz et aI.,
1998,1999b).
Conversion electron M6ssbauer spectroscopy (CEMS) was used to study
the surface crystallization of amorphous alloys, which serve as precursors for
the formation of nanocrystalline phase, and clear differences between the
surface and bulk crystallization of various amorphous alloys were observed
(Kopcewicz and Grabias, 1996).
This chapter will be devoted to the characterization of the structure and
magnetic properties of soft magnetic nanocrystalline alloys by using the
M6ssbauer spectroscopy. Various techniques used in the M6ssbauer
spectroscopy, such as the conventional measurements in transmission
geometry, the rf-M6ssbauer technique in which the effects induced by an
external rf magnetic field in ferromagnetic materials are employed, and the
conversion electron M6ssbauer spectroscopy, which probes a near surface
region of the sample about 120 nm thick, thus allowing the study of the surface
effects, will be introduced. The power and limitations of the M6ssbauer
spectroscopy in the studies of nanocrystalline alloys will be discussed.
M6ssbauer spectroscopy may provide a unique set of information on the
structure and magnetism of nanocrystalline alloys unavailable by conventional
methods, particularly when combined with other structural probes such as
XRD, SAXS and TEM. It wi II be shown how the M6ssbauer results complement
and extend our understanding of the properties of this new and important class
of magnetic materials.
Mossbauer Spectroscopy Characterization of Soft Magnetic... 155

5.2 Mossbauer Spectroscopy

5.2. 1 Principles of the Mossbauer Effect


The phenomenon of the emission and resonant absorption and scattering of
nuclear gamma rays without energy loss due to recoil is commonly known as
the "M6ssbauer effect", in recognition of Rudolf L. M6ssbauer who discovered
the effect in 1958 (M6ssbauer, 1958a, 1958b). The unique feature of this
recoil-free nuclear gamma resonance is the unrivaled accuracy in measuring
the changes of the energy of nuclear gamma radiation related to the fact that
the recoil-free gamma radiation has a natural linewidth. The excellent energy
resolution allows us to study the hyperfine interactions of nuclei with the
electronic charge and the effective magnetic fields and electric field gradients
at the nuclear site. The M6ssbauer effect was observed for 109 transitions in
90 isotopes of 47 elements. However, by far the most important isotope is
57 Fe, followed by 119 Sn. Since its discovery the M6ssbauer effect was
successfully applied in many branches of natural science: physics, chemistry,
materials science, biology, mineralogy, etc.
This section serves as a basic introduction to the fascinating field of
M6ssbauer spectroscopy. The simple qualitative presentation of the principles
of the M6ssbauer effect is given. The conditions for the occurrence of the
M6ssbauer effect and its basic features, such as the parameters of the
resonance line, probability of the effect, M6ssbauer isotopes and experimental
methodology are described. The hyperfine interactions leading to the isomer
shift, quadrupole splitting and magnetic hyperfine splitting as well as their
combination are introduced. It is shown how to derive information concerning
the properties of sol ids from the M6ssbauer spectra.
The way to the discovery of the recoil-free nuclear gamma resonance
commonly known as the "M6ssbauer effect" is led through conventional
investigations of the resonant absorption of nuclear gamma radiation. In the
pre-M6ssbauer period the emission and absorption of nuclear gamma radiation
was treated in a classical way. The energy and momentum conservation in the
process of emission and absorption of gamma radiation was considered for a
free nucleus. The forces binding the nucleus in a solid were ignored. Thus, a
nuclear transition with energy Eo was associated with the recoil of the nucleus
which decreased the energy of the emitted gamma radiation. The recoil
energy E R being large as compared with the width of the emission line is what,
together with the shift of the emission and absorption lines due to recoil,
causes the result that the emission and absorption lines usually do not overlap
156 Michat Kopcewicz

and the resonance does not occur. In the pre-M6ssbauer experiments it was
possible to observe the gamma resonance effect either by compensating the
recoil energy or by thermal broadening of the emission and absorption lines. In
the nineteen fifties R. L. M6ssbauer studied the conventional gamma resonance
effect for 1911r and observed an unexpected increase of the resonance effect at
low temperature which was contradictory to the classic model. R. L.
M6ssbauer interpreted this effect in terms of the Lamb model of capture of
neutrons by atoms in the crystal (Lamb, 1939). The basis of this discovery
constitutes grounds for a fundamentally new concept of the recoil-free emission
of nuclear gamma radiation from nuclei bound in a solid and its resonant recoil-
free absorption introduced by R. L. M6ssbauer in 1958 (M6ssbauer, 1958a,
1958b). The 1911r nuclide did not make this effect famous. A breakthrough was
made when the recoi I-free resonant gamma absorption for 57 Fe was observed
(Schiffer and Marshall, 1959; Hanna et a!., 1960a). It was recognized that
the unique properties of the M6ssbauer effect, such as a very narrow
resonance Iine and excellent energy resolution, allow the study of the hyperfine
interactions. In 1960 the isomer shift, the quadrupole interaction and magnetic
hyperfine structure in ex-Fez 0 3 (Kistner and Sunyar, 1960) and in ferromagnetic
ex-iron (Hanna et ai., 1960b) were observed. In 1961, R. L. M6ssbauer was
awarded the Nobel Prize.
Since its discovery the M6ssbauer effect was applied in many disciplines
of natural science: physics, chemistry, biology, mineralogy. Now it is one of
the most sophisticated experimental methods which allows us to measure with
unrivaled accuracy the changes of energy of gamma radiation.
In order to understand the difference between the recoilless y-emission
from the nucleus bound in a solid (M6ssbauer effect) and the y-emission from a
free nucleus, let us consider the process of emission of a gamma ray from a
free nucleus at rest. Let Eo be the energy of the excited state of the nucleus,
which de-excites to the ground state by emitting a gamma ray. The gamma
emission event is associated with the recoil of the nucleus (to conserve
momentum), so the energy of the gamma ray is Ey = Eo - E R From energy
and momentum conservation the recoil energy of the nucleus can be calculated
as
(5.1)

where M is the mass of the nucleus and p is the momentum transferred to the
nucleus equivalent to the momentum of the gamma photon. Similarly, when a
gamma photon is absorbed it loses the energy E R due to the recoil imparted to
the absorbing nucleus.
In order to find whether the resonant gamma ray absorption will occur the
recoil energy must be compared to the linewidth, r, of the emission
(absorption) line, which is related to the mean lifetime T of the excited state
(5.2)
Mossbauer Spectroscopy Characterization of Soft Magnetic... 157

where T is the mean lifetime of the excited state of the nucleus, and 1'i =2~' h
is Planck's constant. As can be seen from Eq. (5. 1), the recoi I energy
depends strongly on the energy of the gamma rays. For example, for a nucleus
with M= 100 and E y """E o """l x 10 4 eV the recoil energy E R """5 x 1O- 4 eV.
However, if the energy of the gamma rays is 5 x 10 4 eV, then the recoil energy
E R increases already to about 1. 3 x 10- 2 eV. Since the typical natural
linewidth r is of the order of 10- 8 eV for T""" 100 ns, the emission and
absorption lines will be separated by 2E R r and resonance will not occur. In
order to observe the resonant absorption the emission and absorption lines
must at least partially overlap.
Various methods have been used to compensate the recoil in order to
make observation of the gamma resonance possible:
( 1) Line broadening due to thermal motion of emitting and absorbing
nuclei at elevated temperature;
(2) The Doppler shift associated with the motion of the emitting nuclei
(source) toward the absorber with a sufficiently high velocity (usually of the
order of 10 4 cm/s);
(3) Employing the recoil from the preceding nuclear decay or reaction.
In all these techniques the recoil energy is compensated, in clear
distinction to the Mossbauer effect discussed below in which the recoil energy
is eliminated and compensation is not required.
In 1958 Rudolf L. Mossbauer discovered that when the emitting or
absorbing nuclei are not free but bound in a solid, a certain fraction of such
events may occur with negligible energy loss due to recoil. R. L. Mossbauer
made this discovery when studying the scattering of 129 keV gamma rays from
1911r by Ir and Pt at low temperatures (Mossbauer, 1958a, 1958b). In
distinction from the existing models discussed above, Mossbauer observed an
unexpected increase in the scattering when the temperature was decreased.
The new idea of recoil-free gamma emission and absorption can be understood
in terms of a simple phenomenological model. It can be shown that the mean
energy transferred to the solid due to emission of a gamma quantum from a
nucleus bound in the solid is equal to the recoil energy of the free nucleus E R
Three cases can be distinguished:
(1) If the free-atom recoil energy E R is larger than the binding energy of
the atom in the solid (which typically is of the order of 10 eV), then the
emitting atom (nucleus) will be displaced from its lattice site and the situation
will be similar to emission from the free atom. Such a situation is typical for
gamma ray energy of the order of 1 MeV or more. In such a case the nuclear
gamma resonance for lines with natural width will not occur;
(2) If the free-atom recoil energy E R is smaller than the binding energy
but larger than the characteristic energy of lattice vibrations (phonon energy)
defined in terms of the Debye frequency (WD) or Einstein frequency (WE)' the
atom will remain in its lattice position but will dissipate the recoil energy by
158 Michat Kopcewicz

creation of phonons. This situation is typical for gamma ray energies of several
hundred keV. In this case the energy of the emitted gamma ray will be smaller
than the resonance energy by nn Wo or nn WE' and resonance will not occur;
(3) If the recoil energy E R is smaller than the phonon energy n WE' a new
effect arises, because the solid as a quantum system cannot be excited in an
arbitrary way. The recoil energy is not dissipated either by displacing the
nucleus from its lattice site or by heating the lattice (phonons). The nucleus
behaves as if it were rigidly bound to the solid and the recoil is taken by the
entire solid. Such a situation may occur for low gamma ray energies of about
10-150 keV. In this case in Eq. (5.1), M in the denominator represents the
mass of the entire solid making E R negligible as compared with r. When this
occurs in the source and the absorber, then the conditions for nuclear gamma
resonance are fulfilled and a large resonance for emission and absorption lines
with natural width is observed. The recoil-free (zero phonon) emission,
absorption and scattering of nuclear gamma radiation is called the M6ssbauer
effect.
A general expression for the probability of the M6ssbauer effect, i. e., the
fraction of the zero phonon (recoil-free) processes, is given by the same
Debye-Waller (W) factor which is applicable to the scattering of X-rays or
neutrons by atoms:
f = exp(- 2W) = exp(- 41T 2
<X 2 >/A 2 ) = exp[ - 2
K <X
2
>J (5.3)
where i\ is the wavelength of the gamma quantum:
K = 21T/A = E/hc,
and <x 2 > is the component of the mean square vibrational amplitude of the
emitting nucleus in the direction of the gamma ray. A large probability for the
M6ssbauer effect (f close to 1) occurs when K 2 <x 2 > 1, i. e., when the rms
displacement of the nucleus is small as compared to the wavelength i\. This is
why the M6ssbauer effect is not observed in gases and nonviscous liquids,
where <x 2 > is not limited.
Since < x 2 > depends on temperature, the study of the temperature
dependence of the M6ssbauer effect provides useful information concerning
lattice dynamics. The probability of the M6ssbauer effect can be calculated in
terms of the Debye model of lattice dynamics. Only <x 2 > must be calculated
because K is constant for a given gamma transition. The Debye model yields
the following expression for the probability of the recoil-free fraction,
(Greenwood and Gibb, 1971)

6E (1
2 o
T f0 /T xdx )J = exp( -
f = exp [ - k B ~o 4 + e~ 0 eX _ 1 2 W) (5.4)

where eo is the Debye temperature, k B is the Bolzmann constant , T is the


temperature and
x = nw/k B T.
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 159

The recoil-free fraction f can be evaluated in the high and low


temperature limits with respect to 0 D . At low temperatures (T 0 D ) ,
2 2
_
f - exp - k B 0
[ ER
D
( 23 + 1T0~T ) ]
. (5.5)

At absolute zero (T = 0 K) ,
f = exp( - 3E R /2k B 0 D ). (5.6)
In the high temperature range (T 0 D!2) ,
f = exp(-6ERT/kB0~). (5.7)

Equation (5.7) yields the linear dependence of the Debye-Waller factor W on


T at high temperatures. This is true only when the harmonic approximation
(Debye model) holds. In many experimental cases, when anharmonicity plays
a significant role a deviation from the proportionality of W to T is observed.
Concluding, the probability of the Mossbauer effect depends on three
factors:
( 1) Free atom recoil energy, related to the gamma transition energy
(Eq. (5.1));
(2) The properties of the solid (i. e., the Debye temperature);
(3) The temperature of the solid (T).
A large probability is observed for low energy gamma transitions in a solid
with high Debye temperature when the measurements are performed at low
temperature.
In the resonance experiment an emission line of the width r s is moved
over an absorption line of the width r . The experimentally observed shape of
the resonance line is:

aexp(E) = [00 w(e)a(E - e)de (5.8)

where

(5.9)

and
(r./2)2
a(E)=ao 2 (5.10)
(E - E y ) + (r. /2)
2'

In the thin absorber approximation limit,

(
E) (rs + r.)2/4 (5.11)
a exp = ao (E - E y )2 + (r s + r.)24
If rs~r.~r, where r is the natural linewidth, then
r2
a exp ( E) ~ a 0 (E _ E y ) 2 + r2 . (5.12)
160 Michat Kopcewicz

Since the theoretical line shape is well known (Eq. (5. 12)), the Mossbauer
effect allows us to measure the changes of the gamma radiation energy of the order
of a fraction of the natural Iinewidth I. The energy resolution, defined as
(5. 13)
is of the order of 10- 15 - 10- 12 depending on the Mossbauer isotope. The
Mossbauer effect provides the most accurate method for measuring the
changes of the energy of electromagnetic radiation. Due to this feature the
Mossbauer effect is the most powerful technique for studying hyperfine
interactions.
The maximum resonance absorption occurs when the energies E y in the
source (E~) and absorber (E~) coincide. The effective E y value can be
changed by moving the source with respect to the absorber with velocity v,
e. g., by using an externally appl ied Doppler effect. The change of the energy
E y due to such motion is
/lE = (v/c)E y . (5. 14)
Usually absorption experiments are performed, so the maximum
resonance absorption correspohds to the minimum count rate at the velocity
v" for which E~ = ( E~). At any higher or lower velocity the resonance
absorption decreases until it finally vanishes at velocity far from v,. The shape
of the absorption line is Lorentzian (Eq. (5. 11)) with the full width at half
maximum l r = Is + I., and the count rate at v gives the nonresonant
00

background. The nuclei in the absorber excited due to resonance absorption


re-emit gamma rays within T""'" 10- 7 s. However, when the internal conversion
coefficient is high, relatively few gamma rays wi II be re-em itted. The
re-emission process is not directional but occurs over the 41l" solid angle. Thus
the number of counts in the detector corresponding to re-emitted gamma rays
in a collimated, strongly directional transmission experiment is usually
negligible.
Concluding, the parameters of a simple single-line Mossbauer spectrum
are as follows:
( 1) position of the resonance v,;
(2) full width at half maximum of the resonance line l r = Is + 1."""'21;
(3) depth of the resonance line expressed as

- N, - Noo
Or - Noo '

where N, and N oo is the count rate at resonance and far from resonance (at
v00 respectively;
),

(4) nonresonant background count rate at v 00 ;

(5) area of the resonance line:

(5. 15)
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 161

In order to observe the Mbssbauer effect, a proper isotope must be found


and favorable conditions for the solid matrix in which the isotope will be bound
must be fulfilled. The Mbssbauer nucleus should fulfil the following basic
requirements:
(1) low gamma transition energy (typically 10 - 150 keV) between the
first excited and the stable ground state, and nucleus of the mass M> 50 to
minimize the recoil energy E R ;
(2) mean lifetime of the excited level T~1O-9 -1O- 7s; a longer lifetime
results in a very narrow linewidth that makes the resonance difficult to observe
experimentally, while a shorter lifetime results in a wide resonance line with
poor resolution;
(3) the preferable multipolarity of the gamma transition M1 or E2;
(4) isotope production in the excited state as a product of natural
radioactive decay of the parent isotope; the lifetime of the parent isotope
should be reasonable, preferably of the order of one year or more;
(5) large maximum resonance cross-section ao.
By far the best combination of the requirements listed above is found in
57 Fe which is the most convenient and most popular Mbssbauer isotope.
However, the Mbssbauer effect was observed for 109 transitions in 90 isotopes
of 47 elements.
We will limit our discussion to the 57Fe Mbssbauer spectroscopy because
most of Mbssbauer investigations involve the 14.4 keV transition in 57 Fe. It is
fortunate that the element of great commercial importance and that is so
abundant in nature which forms a great variety of chemical compounds and
alloys has an isotope with an almost ideal combination of properties for
Mbssbauer spectroscopy. The gamma transition for which the Mbssbauer
effect is observed has very low energy (14. 39 keV), and hence, the recoil
energy is small (1. 95 X 10- 3 eV); an almost purely magnetic dipole (M1)
transition (E2 admixture is negligible) occurs between the first excited state

(I. = ~ ) and the ground state (/ g = ~ ). The lifetime (TI/2) of this level is
about 100 ns, hence the natural linewidth is only 0.45 X 10- 8 eV and produces
the Iinewidth of the Mbssbauer Iine of 2 r = O. 194 mm/s. The small amount of
57Fe in natural iron (2.19%) is compensated by the large cross-section for
resonant absorption, ao~257 x 10- 20 cm 2 . The 14.4 keV transition has an
internal conversion factor a ~ 10 which allows us to use the conversion
electrons for detecting the Mbssbauer transition.
57 Fe nucleus is produced by f3 + (electron capture) decay from the parent
57 Co isotope whose half Iifetime is 270 days. 57 Co decay populates the second
excited state in 57 Fe of the energy of 136 keV which decays in 90 % through a
cascade to the 14. 4 keV level and then to the ground state, and in 10%
directly to the ground state. The Mbssbauer effect was also detected for the
136 keV transition but its importance for the applications of the Mbssbauer
spectroscopy is negligible. The recoil energy being small, the probability of
162 Michat Kopcewicz

the recoil-free emission and absorption of 14.4 keV gamma rays is large even
at high temperatures (well above room temperature). Thus the Mbssbauer
effect can be employed to study hyperfine interactions in a wide temperature
range.
By using the Mbssbauer effect it is possible to measure the changes of the
energy of gamma radiation with astonishing accuracy. The energy resolution of
the 57 Fe resonance is ,.../ E y """'3 x 10 -13 .
The most typical Mbssbauer experiment involves a radioactive source
containing the Mbssbauer isotope in an excited state and an absorber
consisting of the material to be investigated which contains the same isotope in
the ground state. Here we will present the most common case of the
experiments with the use of the 57 Fe isotope. In transmission geometry the
gamma rays emitted from the source pass through the absorber in which
resonant absorption occurs and then they are detected by a suitable detector,
usually a proportional counter or scintillation Nal (T1) detector. In order to
investigate the energy levels of the Mbssbauer nucleus in the absorber which
may be shifted with respect to the energy levels of the same isotope in the
source material or split due to hyperfine interactions, it is necessary to
modulate the energy of the emitted gamma radiation so that it can match its
energy to the resonance. The most common method of energy modulation is
the Doppler modulation consisting in a relative movement of the source with
respect to the absorber. The energy change 8E is given by Eq. (5. 14). For
the 14. 4 keV transition in 57 Fe the velocity of 1 mm/s corresponds to 4. 8 X
10- 8 eV or 11.6 MHz. The motion of the source is usually oscillatory to provide
an energy scan.
The typical velocity transducer is an electromechanical device consisting
of two coils fixed to the rod supported by springs and placed in the static
magnetic field (double loudspeaker system). The drive signal delivered from
the drive generator to one of the coils induces the motion of the rod, to which
the source is attached. Due to this motion the pick-up signal is induced in the
pick-up coil which is used for correcting the shape of the drive signal via a
negative feed-back loop. The shape of the drive signal is such that the motion
occurs with constant acceleration (linear velocity scale). Sometimes a
sinusoidal drive signal is used.
The detector, preamplifier, amplifier and single channel analyzer (SCA)
form a conventional gamma spectrometer. All pulses from the detector,
regardless of their energy, are ampl ified. The SCA is used for selecting from
the whole energy spectrum only the pulses corresponding to the Mbssbauer
gamma radiation. These pulses are delivered to the acquisition system
working in the multiscaler mode (MSC). The address of the MSC is triggered
by a starting pulse from the drive generator, so the counting is synchronized
with the motion of the source in such a way that each channel in the multiscaler
stores the counts corresponding to a given velocity. In this way a Mbssbauer
spectrum, i. e., counting rate vs. velocity is obtained.
Mossbauer Spectroscopy Characterization of Soft Magnetic... 163

The experiments are performed not only in transmission geometry (which


is most common) but also in a scattering geometry. The scattering geometry is
useful when a thin absorber cannot be prepared and it allows examination of
the thin surface layer of the bulk materials. In the scattering geometry the
Mossbauer effect can be monitored either by detecting the Mossbauer gamma
radiation or conversion electrons (conversion electron Mossbauer
spectroscopy- -CEMS), or else by X-rays (conversion X-ray Mossbauer
spectroscopy - CXMS) associated with the gamma absorption process.
The CEMS technique is of particular interest for the study of
nanocrystalline alloys. The decay of resonantly excited 14.4 keV level in 57Fe
is a complex process. The de-excitation may occur either directly via the
emission of a gamma quantum or via internal conversion as a result of which
conversion electrons from K, L or M electronic shells are emitted. The most
numerous are 7. 3 keV K-conversion electrons (probability of occurrence is
81 % ), followed by 13. 6 keV L-conversion electrons (9 % ). Emission of
K-conversion electron creates a hole in the K-shell as a result of which 5.6 keV
KLL Auger electrons (53 %) and 6.4 keV K a X-rays (27 %) are emitted. In the
most common CEMS technique all emitted electrons are registered without
energy discrimination. The electrons are usually detected by the gas flow
proportional counter with He-6 % CH 4 working gas. The sample is placed inside
the counter and the Mossbauer measurements are carried out in the
backscatter geometry. Since the counter registers only the electrons emitted
from the sample the depth of the surface region probed by this technique is
related to the maximum energy of electrons. In the case of iron sample about
85 % of 7.3 keV K-electrons comes from the depth of about 120 nm, and 96 %
from less than 300 nm. Thus the CEMS technique is very useful for probing the
near surface regions of materials.
Much more sophisticated CEMS technique makes use of the electron
energy analysis. Setting an energy window by means of electron energy
analyzer, one can obtain information from a well selected depth of the sample.
This technique, called a depth selective conversion electron Mossbauer
spectroscopy (DCEMS), is much more difficult and time consuming, compared
to conventional CEMS without energy selection.
Similar to CEMS the technique is the conversion X-ray Mossbauer
spectroscopy (CXMS) in which the 6 4 keV K a X-rays, originating from the
hole in the K-shell formed due to emission of the K-conversion electrons, are
detected by a similar gas flow counter with Ar-working gas or conventional
proportional counter with good efficiency for the low energy X-ray or gamma
radiation. The probing depth for CXMS is about lO\Jm.
The Mossbauer spectroseopy provides all information regarding via the
hyperfine interactions.
164 Michat Kopcewicz

5.2.2 Hyperfine Interactions

The great accuracy of the Mossbauer effect of measuring the changes of the
energy of electromagnetic (nuclear gamma) radiation is the most important
feature of this technique. The energy resolution for the 57 Fe-Mossbauer effect,
related to the linewidth of the resonance line, is of the order of 10- 13 , which
means that gamma radiation energy changes of the order of 10- 8 eV can be
measured. Therefore the Mossbauer effect allows us to observe directly the
shifts and spl ittings of the Mossbauer Iines resulting from the interactions of the
nucleus with electrons. The hyperfine coupling mechanisms yield information
regarding electron and spin density distributions, thus providing information
concerning chemical, structural and magnetic properties of solids.
The hyperfine interaction Hamiltonian for the atom contains terms related
to interactions between the nucleus and its environment (electrons)
H = Ec + M1 + E2 + ... (5. 16)

where E c refers to the Coulomb interaction between the nucleus and electrons
at the nuclear site, M 1 is the interaction between the nuclear magnetic dipole
moment and the effective magnetic field at the nucleus (magnetic dipole
hyperfine interaction), and E 2 is the interaction between the nuclear
quadrupole moment and the electric field gradient at the nucleus (electric
quadrupole interaction). These are the most important interactions, which
determine the shape of the Mossbauer spectra. The term E c describes the
isomer shift, M 1 is the nuclear Zeeman effect which is responsible for the
magnetic hyperfine structure, and E 2 causes the quadrupole spl itting.
Interactions of higher order (M 3 , E 4' etc.) can be neglected because their
energies are by several orders of magnitude smaller than Eo, M 1 and E 2 , and
the electric dipole interaction (E 1) is parity forbidden.

5. 2. 2. 1 Isomer Shift
The absorption line in the Mossbauer spectrum is shifted as a result of electric
Coulomb interaction between the nuclear charge distribution over a nuclear
volume and the electronic charge density at the nucleus. The nucleus is
considered as a uniformly charged sphere, the radius of which is different in
the ground and the excited state. The elctronic charge density at the
Mossbauer nucleus depends on chemical properties and differs for various
materials. Since in the Mossbauer experiments the source and the absorber
are involved, for which the chemical environments of resonant nuclei are
usually different, the shifts of nuclear levels are different. That results in the
shift of the Mossbauer spectral lines. The isomer shift is measured with
respect to the source material or to the "standard" material which determines
the reference point on the scale of isomer shifts. For 57 Fe-Mossabuer
Mossbauer Spectroscopy Characterization of Soft Magnetic... 165

spectroscopy such a reference standard with respect to which the isomer shifts
are reported is the center of the magnetically split (see below) spectrum of
ferromagnetic ex-Fe at room temperature. It is not possible to separate in a
single experiment the contributions to the observed shift resulting from a
nuclear part and the electronic part of the interaction. Usually the nuclear part,
constant for a given gamma transition, is established first, or measured by
another method, and the isomer shift is used to determine the electronic term.
The isomer shift provides valuable chemical information. It reflects the
changes in the electron density at the nucleus due to changes in the valence
orbital population of the M6ssbauer atom. The electron density is related to
the type of chemical bonding, covalency effect, oxidation and reduction
processes, and differences in the electronegativity of ligands coordinated to
the Mossbauer atom. The study of a single compound is of little value for
obtaining useful information unless data for other compounds are available for
comparison. Only then is it possible to correlate isomer shift data with the
oxidation and spin state of the atom in a complex compound, the coordination
between the Mossbauer atom and bonded ligands, etc. The richest data have
been collected for iron compounds. The isomer shift is sensitive enough to the
spin state of iron in a complex compound to allow distinguishing between high
spin and low spin divalent and trivalent iron. The isomer shift data alone make
the distinction between low spin iron ( II ) and iron ( ill) difficult, but with the
help of quadrupole splitting (which is usually substantially larger for iron( ill)
compounds) they can be readily identified. High spin iron compounds can be
distinguished by isomer shift data alone.
The line position in the spectrum is determined not only by the isomer shift
but also by a relativistic temperature-dependent contribution related to thermal
vibration of the nuclei (second order Doppler shift). This shift is given by the
mean square velocity of the resonant nuclei related to the thermal motion and
can be calculated in terms of the lattice dynamics model (Einstein or Debye
model). Therefore the temperature difference between the source and the
absorber should be taken into account when quoting and comparing the isomer
shifts.

5. 2. 2. 2 Quadrupole Splitting
The nuclei in states with spins larger than 1/2 have non-spherical electric
charge distributions which are characterized by a nuclear quadrupole moment.
When a non-spherical nucleus experiences an asymmetric electric field,
described by the electric field gradient (EFG), the electric quadrupole
hyperfine interaction occurs which leads to a splitting of the nuclear levels
corresponding to different alignments of the quadrupole moment with respect to
the electric field gradient. If the 57 Fe nucleus encounters the non-zero EFG than
the excited state (J =3/2) splits into two sublevels (m/ = 1/2 and 3/2).
The ground state (I = 1/2) remains unsplit because the nucleus with spin 1=
1/2 is spherical and does not have a quadrupole moment. The spectrum
166 Michat Kopcewicz

consists now of two lines, related to the gamma transitions between the ground
state and two sublevels of the excited state, separated by the "quadrupole
splitting". As for the case of isomer shift, the quadrupole splitting depends on
the nuclear parameter (nuclear quadrupole moment) and the electronic
parameter (EFG at the nucleus site) which for 57 Fe cannot be separated in a
single experiment. Usually the nuclear quadrupole moment, constant for a
given nuclear level, is established first, and the details of the electric field
gradient are of primary interest.
The quadrupole interaction provides information about the electric field
gradient at the nucleus site, which can originate either from the valence
electrons of the M6ssbauer atom, when it is associated with asymmetry in the
electronic structure, or from a non-spherical charge distribution in the ligand
sphere and/or lattice surrounding with symmetry lower than cubic. Molecular
orbitals can also contribute to EFG. The effects of these contributions at the
nucleus site are modified by the polarization of the core electrons of the
M6ssbauer atom which may reduce or enhance the EFG. The quadrupole
splitting observed in the M6ssbauer spectra of a given solid reflects the
symmetry of the bonding environment and the local structure in the vicinity of
the M6ssbauer atom. Quadrupole splitting data are especially useful when
combined with isomer shift data. They yield chemical information concerning
electronic population of various orbitals, ligand structure and structural
information on the local atomic arrangement both in crystalline and amorphous
sol ids. Quadrupole spl itting is a particularly sensitive probe of short range
order in metallic glasses.
S. 2. 2. 3 Magnetic Splitting
When a nucleus experiences a magnetic field the magnetic hyperfine
interaction occurs which couples the nuclear magnetic moments of the ground
and excited states to the magnetic field acting at the nucleus site. This
interaction raises the degeneracy of the nuclear states with spin 1>0 and splits
them into 21 + 1 substates. In the case of 57Fe the ground state (/ = 1/2) splits
into two substates and the excited state (I = 3/2) splits into four substates.
Since only transitions with l:!.m = 0 and 1 are allowed, there are six possible
transitions and hence the M6ssbauer spectrum splits into six lines (Zeeman
sextet). Since the magnetic splitting of the spectrum is directly proportional to
the magnetic field at the nucleus site, the M6ssbauer spectroscopy permits the
determination of this field. The probabilities of transition between the nuclear
substates influence the intensities of the lines in the M6ssbauer spectrum
therefore allowing the determination of the relative orientation of the magnetic
field at the nucleus and the direction of gamma rays.
The magnetic hyperfine interaction Hamiltonian contains a nuclear
parameter (magnetic moment of the nuclear ground and excited states, IJ 9 and
lJe' respectively) and an atomic parameter (magnetic field at the nucleus, the
hyperfine field, H hf ) which cannot be separated experimentally. The situation
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 167

is more favorable here than in the case of the isomer shift and quadrupole
splitting because both nuclear states are split and it is possible to determine in
one experiment two of the three parameters which describe the interaction:
IJ g , lJe and H hf Usually the magnetic moment of the ground state, IJg' is
known, or determined from some place other than M6ssbauer experiments
(e. g., microwave resonance, atomic beam experiments), so lJe and H hf can
be found. One can also apply a sufficiently strong external magnetic field and
observe the resulting splitting, whereas it is not possible to apply a sufficiently
strong external electric field gradient in the case of quadrupole interaction.
The magnetic hyperfine splitting in the spectra yields an effective
magnetic field, at the site of the nucleus, originating from various sources
which are difficult to separate. The total magnetic field experienced by the
nucleus is a vector sum of the various contributions. The largest contribution to
hyperfine magnetic field results from the Fermi contact interaction. Its origin is
either related to the intrinsic impairing of s-electrons, or indirectly results from
polarization effects on the filled s-orbitals due to unpaired d- or f-orbitals.
Other contributions are related to the non-zero orbital magnetic moment of the
atom and the dipolar interaction of the nucleus with the spin of the atom.
The magnetic hyperfine interaction yields information on the magnetic
properties of solids such as magnetic ordering and structure of magnetically
ordered systems, the nature of magnetic interaction (ferro- or
antiferromagnetic), the magnitude of the magnetic moment on particular atom.
This information results from studying the magnetic hyperfine fields at the
nucleus site. However, in order to compare such results with those obtained
by conventional methods (e. g., magnetization measurements) phenomenological
formulas have to be used to convert the information at the site of the nucleus to
those at the site of the atom. Thus the estimation of the magnetic moment of
the atom is indirect. The temperature dependence of the hyperfine field is
fortunately the same as that of magnetization, hence the onset of the magnetic
ordering and magnetic phase transitions can be studied directly by the
M6ssbauer technique. Also the shapes of the hyperfine field H hI ( n and
magnetization M ( n curves are identical and allow the study of deviation of
experimental H hf ( T) curves from the Brillouin function. Combined electric
quadrupole and magnetic dipole interactions allow the study of the direction of
the magnetic field with respect to the crystallographic axes. The application of
the external magnetic field to a system with no unpaired spins, which has no
magnetic hyperfine field, leads to a magnetic splitting in addition to the
quadrupole splitting and provides information on the geometry of the electric
field gradient (EFG) at the nucleus.
From the complex spectra revealing a poorly resolved magnetic hyperfine
structure of ferromagnetic amorphous alloys, it is possible to extract the
distributions of hyperfine fields P ( H) which provide valuable information on
the magnetic properties and structure of such materials.
The M6ssbauer effect allows us to study not only static magnetic fields but
168 Michat Kopcewicz

also time dependent relaxation effects. Every spectroscopy has a


characteristic observation time of interaction of the radiation with the solid.
The observation time of the Mossbauer spectroscopy corresponds to the mean
lifetime (TN) of the nuclear resonant level. For hyperfine interaction
measurements, with the characteristic energy I:::.E, it corresponds to the
Larmor precession period TL =n / I:::.E. Times TN and TL are typically of the
order of 10- 8 - 10- 7 s. In the presence of time dependent changes in the
environment of the resonant nucleus the experimental observation (shape of
the Mossbauer spectrum and the hyperfine parameters) depends on the
relative order of magnitude of the observation time (TN or TL) as compared
with the residence time of the fluctuating environment, TR (relaxation time).
When the observation time is much longer than the relaxation time (TN. TL
TR' fast relaxation) then a fully time-averaged situation occurs. If the opposite

relation occurs (TN. TL TR' slow relaxation) then the quasi-static situation is
observed. For intermediate relaxation rates we get complex spectra.
The time dependence can be introduced into the hyperfine interactions by
using an external high frequency magnetic field. When the frequency of such a
field is higher than the Larmor frequency, then the magnetic hyperfine splitting
collapses and the Mossbauer spectrum consists of a single line or a quadrupole
doublet instead of a Zeeman sextet. The shape of the spectra depends on the
relation between the frequency of the field and the Larmor frequency.

5.2.3 The rf-Mossbauer Technique


The rf-Mossbauer technique is an unconventional specialized technique which
combines the Mossbauer effect with the phenomena induced in ferromagnetic
materials by an external radio-frequency (rf) magnetic field: the rf-collapse
and rf-sidebands.
( 1) The rf-collapse effect occurs when the magnetic rf field forces the fast
reversal of sample magnetization. This results in the oscillations of the
hyperfine field at the Mossbauer nuclei in response to the applied rf field
(Pfeiffer, 1972; Albanese et aI., 1971; Kopcewicz, 1978).
The rf-collapse effect is interpreted in terms of the extreme ferromagnetic
enhancement of the rf field (Pfeiffer, 1972). Let us consider first the static
situation when an external magnetic field is applied to a ferromagnetic
material. When this field is larger than the anisotropy field, Han, saturation
magnetization is reached. The magnetic field at the 57 Fe Mossbauer atom is
then of the order of 104 Oe and directed parallel to the external field. This field
causes the polarization of the iron core electrons and results in a hyperfine field
of about 300 kOe at the sites of the Mossbauer nuclei directed antiparallel to
the magnetization vector. If the material is magnetically soft, a small external
field of several Oe is able to control the direction of a very large hyperfine
field. If the direction of the external field is reversed, then the hyperfine field
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 169

is rotated to the same extent. When a radio-frequency magnetic field is used


instead of the static field, the magnetic hyperfine field will be forced to
oscillate is response to the external rf field, and, as a result, the hyperfine
field at the Mbssbauer nuclei will be reduced. In the extreme case, when the
frequency of the magnetization reversal, and, hence of the reversal of the
hyperfine field is higher than the Larmor precession frequency of the magnetic
moment of an 57 Fe Mbssbauer nucleus about the local hyperfine field, the
magnetic hyperfine field experienced by the Mbssbauer nuclei will average to
zero. As a result a collapse of the entire Zeeman sextet to a single line or a
quadrupole doublet is observed in the Mbssbauer spectrum in spite of the
sample being in the ferromagnetic state. The apparent magnetic splitting will
disappear because of the fast reversal of magnetization when all spins are
coupled and reversed together.
The complete collapse of the magnetic hyperfine splitting occurs when two
conditions are fulfilled. First, the frequency of the external driving rf field must
be considerably larger than the Larmor frequency. Second, the switching time
of magnetization reversal must be comparable to, or shorter than, the period
of the rf field applied. The switching time depends on the properties of the
material and decreases with increasing intensity of the driving rf field
(Pfeiffer, 1972). Hence, the shape of the rf-collapsed Mbssbauer spectrum
depends on both the frequency and intensity of the applied rf field.
Because the rf-collapse effect allows us to average the magnetic hyperfine
field to zero in the ferromagnetic state, the magnetic hyperfine structure
disappears in the Mbssbauer spectrum, so one can observe directly the
quadrupole interaction in the ferromagnetic state. Thus, the rf-collapse gives
us a unique opportunity to separate the magnetic dipole and the electric
quadrupole hyperfine interactions in the ferromagnetic state.
Since the rf-collapse effect strongly depends on the switching time of
magnetization reversal, which, in turn, is very sensitive to the relation
between the driving field intensity and the local magnetic anisotropy field,
information regarding the anisotropy fields can be derived from the
dependence of the shape of the rf-collapsed spectra on the rf field intensity.
The importance of both these features of the rf-collapse effect for the
application of the rf-Mbssbauer technique for the study of the structure and
magnetic properties of soft magnetic nanocrystall ine alloys will be
demonstrated in the following sections of this chapter.
(2) The rf-sidebands effect originates from the rf-induced vibrations of
atoms via magneto-acoustic coupling, which is magnetostriction. When the rf
magnetic field is applied to ferromagnetic material additional lines appear in
the Mbssbauer spectrum. These new lines are separated from the originally
observed lines by integral multiples of the frequency of the applied field. An
infinite set of satellite lines (called sidebands) is formed in relation to each
carrier line of the Zeeman sextet (or of the rf-collapsed pattern). The
rf-sidebands effect can be described by a classical frequency modulation model
170 Michat Kopcewicz

which determines the positions of sideband lines and the intensities


(modulation index) of each order of sidebands (Pfeiffer, 1972; Pfeiffer et al.,
1972; Asti et aI., 1969; Kopcewicz et al., 1975).
The modulation index can be calculated in terms of a simple
magnetostriction model under the assumption that the magnetostriction strain is
given by a static model, the radio-frequency component of magnetization is
much smaller than the static magnetization, the static magnetostriction
constant is appropriate at high frequencies, and, as the acoustic vibrations
generated by the rf field propagate in the sample, interaction with grain
boundaries, defects, surfaces, dislocations, etc., develops components in
directions other than their original one. The modulation index resulting from
this model is proportional to the magnetostriction constant. Therefore the
intensities of rf-sidebands are directly related to magnetostriction, thus
allowing us to follow the changes of magnetostriction in the nanocrystalline
alloys related, e. g., to the formation and development of the nanocrystalline
grains in the residual amorphous matrix.
The rf-collapsed effect can be induced only in the soft ferromagnetic
materials because of the condition that the rf field intensity should exceed the
magnetic anisotropy. The rf-sidebands effect can occur only in the
magnetostrictive ferromagnetic materials, i. e., below the Curie point. It
disappears when magnetostriction vanishes.
If the material exposed to the rf field is a soft magnetostrictive
ferromagnet then both rf-induced effects coexist. They may be separated
either by suppressing the rf-collapse (e. g., by inducing external anisotropy
(Kopcewicz et aI., 1985) or by static magnetic field superimposed on the rf
field (Pfeiffer et aI., 1972; Kopcewicz et aI., 1985 or by suppressing the
rf-sidebands (e. g., by covering the sample surface with a damping medium
(Albanese et al., 1972; Kopcewicz et aI., 1975; Kopcewicz and Kotlicki,
1980. Therefore, despite the fact that both effects are induced by the same
rf field, their origin is different: the rf-collapse is related to the fast relaxation
of the hyperfine field resulting from the fast magnetization reversal, and the
rf-sidebands originate from frequency modulation of the M6ssbauer gamma
radiation by the acoustic vibrations induced by the rf field via
magnetostriction. The rf-collapse requires the rf field frequency to be several
times larger than the Larmor frequency, while the rf-sidebands effect occurs at
all frequencies. However, to observe the well-resolved sidebands in the
M6ssbauer spectrum, the rf field frequency should be sufficiently large so that
the sidebands separation from the carrier lines exceeds the linewidth of the
carrier line. The sideband intensity depends on frequency and strongly
decreases for higher frequencies (Pfeiffer, 1972; Pfeiffer et aI., 1972).
Typical rf-M6ssbauer experiment is performed in transmission geometry.
The sample, playing a role of absorber, is placed inside the coil to which the
sinusoidal signal of a given frequency is delivered from the rf power generator.
The rf magnetic field is applied in the plane of the sample; so, it is
Mossbauer Spectroscopy Characterization of Soft Magnetic... 171

perpendicular to the direction of propagation of the M6ssbauer gamma


radiation. Usually, because of the rf heating (eddy currents and hysteresis
heating), the sample must be cooled to keep its temperature well below the
Curie point. In the experiments described in this chapter, the sample holder
was water-cooled. In order to observe a complete rf-collapse of the magnetic
hyperfine structure in the M6ssbauer spectrum of ferromagnetic amorphous or
nanocrystall ine alloys, the rf field frequency of about 60 MHz (about 3 - 4
times larger than Larmor frequency) was sufficient. Typical rf field intensity
was varied between 0 and 20 Oe. A simple rf power generator of about 100 W
was used. In most experiments discussed here, the frequency was kept
constant and the intensity of the rf field was changed by changing the anode
voltage, which controlled the rf current in the coil. The pick-up coil was used
to monitor the rf field intensity. The rf coil with the sample inside has to be
well screened from the rest of the M6ssbauer spectrometer to prevent the rf
signal from affecting the M6ssbauer electronics. The rest of the spectrometer
is conventional.
The M6ssbauer experiments in the rf fields were reviewed by Pfeiffer
(1972), Srivastava (1983), and Kopcewicz (1989, 1991).
The M6ssbauer studies of the rf-induced effects evolved from the
investigations of the origin and basic properties of these effects (Pfeiffer,
1972; Srivastava, 1983) to the specialized rf-M6ssbauer technique
(Kopcewicz, 1987, 1991) which may provide unique information regarding the
structural and magnetic properties of various ferromagnetic materials,
including amorphous and nanocrystalline alloys.

5.3 Experiment

5.3.1 Formation of the Crystalline Nanostructure in


Amorphous Matrix
Amorphous alloys most appropriate for the formation of the nanocrystalline
phase are characterized by two steps in their crystallization behavior, which
appear as two well separated exothermic peaks in the differential scanning
calorimetry (DSC) curve. These steps correspond to two crystallization
temperatures related with the primary and secondary crystallization. The
primary crystallization consists in the formation of crystalline nanograins,
typically 10 - 15 nm in diameter, which precipitate randomly in the residual
amorphous structure. At the secondary crystallization, at temperatures
exceeding the second DSC peak, a complete crystall ization of the residual
amorphous phase occurs, and the crystall ine phase grows in much larger,
172 Michat Kopcewicz

microsize regions. In order to form the nanocrystalline alloy, amorphous


precursor is annealed in protective atmosphere or in vacuum (to avoid
oxidation) at temperatures above the primary crystallization temperature but
below the second crystallization temperature. Therefore the most suitable
amorphous precursors reveal large differences between the temperatures of
the first and the second crystallization steps. Typical time of annealing the
amorphous precursor is usually 1 h. However, in order to study the
crystallization kinetics annealing have been performed as a function of time at
fixed temperature starting from annealing time of the order of 1 s.
Three groups of soft magnetic nanocrystall ine alloys can be distinguished:
(1) the FINEMET-type based on Fe-B-Si amorphous alloys with the
addition of small amount of Cu and Nb. Formation of the nanostructure in
amorphous FeCuNbSiB alloy with 70 % - 80 % of Fe is explained by a
combination of the enhanced nucleation rate of the bcc phase resulting from the
immiscibility of Cu in Fe and a reduced crystal growth rate related to small
diffusivity of Nb in Fe. The nanostructure formed due to annealing consists of
the bcc-Fe(SD phase, tending to the Fe3Si with the D0 3 structure. The first
and best-known alloy from this group, the Fen5 CUI Nb 3Si 135 B9 , was used in
the pioneering work by Yoshizawa et al. (1988) in which formation of the soft
magnetic nanocrystalline alloy was demonstrated. Nanocrystalline FINEMET
alloys show excellent soft magnetic properties, but exhibit, however, lower
saturation magnetic flux density (B s) of about 1. 3 - 1. 5 T, as compared with
Fe-metalloid amorphous alloys because of the decrease of the Fe content by
the addition of Cu and Nb;
(2) The NANOPERM-type based on Fe-M-B (M: Zr, Hf, Nb, Ti, Ta ,
Mo, ... ) alloys with 85 % - 92 % of Fe, with eventual addition of Cu. The
nanocrystalline phase formed by annealing is almost pure bcc-Fe embedded in
the retained amorphous matrix (Suzuki et al., 1990, 1991 a, 1991 b, 1993;
Mak ino et al., 1995, 1997a, 1997b). The B s of such nanocrystall ine alloys is
higher than those of the FINEMET-type alloys, typically 1.5 - 1.7 T;
(3) The HITPERM-type, based on the FeCoZrBCu alloy. The precursor
for the HITPERM nanocrystalline alloys is the amorphous Fe 4 4 C0 44 Zr7 B4 CUI
alloy, which reveals much higher crystallization temperature and Curie point as
compared with NANOPERM alloys (Willard et al., 1998, 1999). The
nanophase formed by annealing is the bcc-FeCo. Because of high stability at
elevated temperatures related to the high crystallization temperature, the
HITPERM alloys are expected to find applications at much higher working
temperatures than NANOPRM and FINEMET alloys.
Mossbauer Spectroscopy Characterization of Soft Magnetic... 173

5.3.2 Primary Characterization of Thermal Stability of Amorphous


Precursor- Formation of Nanostructure
5.3.2.1 Differential Scanning Calorimetry
The crystallization behavior of amorphous alloys, serving as the precursors in
which the nanocrystalline phase can be formed by annealing, is usually studied
first by differentials scanning calorimetry (DSC). The DSC measurements
allow the determination of the type of crystallization (one or many steps) and
the determination of the crystallization temperatures. It should be noticed that
the DSC measurements were routinely performed in a majority of investigations
in which the formation of the nanocrystalline phase embedded in amorphous
matrix was of primary importance. As an example, the DSC results obtained
for the amorphous Fe93-x-yZr78xCuy(x=6, 8 and 12; y=O, 2) (Kopcewicz
et aI., 1996, 1997a,) and FesoM78,2CUl (M: Ti, Ta, Nb, Mo) (Miglierini
et aI., 1999; Kopcewicz et al., 1999a) will be shown.
Amorphous Fe93- x- yZr7 8 x Cuyalloys are crystallized through two steps as
revealed by two exothermic peaks in DSC curves (onset temperatures T x, and
T X2 are indicated in Fig. 5.1) (Kopcewicz et aI., 1995c). The crystallization
temperatures depend on the boron content and for a given boron level
decrease markedly upon the addition of 2 % Cu. The first stage with the
crystallization temperature (T x, ) corresponds to the structural transformation
from the amorphous to the nanoscale bcc phase. The second exothermic peak
( T x 2 ) reflects the transformation of the remaining amorphous phase to a
mixture of FeZr compounds, iron borides and microcrystalline ex-Fe. The
nanocrystalline bcc Fe phase can be formed in the temperature range between
T XI and T X2 ' The 2 % Cu content not only decreases the T XI temperature but
also extends the temperature range between T x , and T X2 (Fig. 5.1).
As can be seen from Table 5.1 (Kopcewicz et aI., 1997a). a decrease of
boron content in the alloys causes the shift of T XI and T X2 towards lower
temperatures and addition of 2 % Cu decreases T XI

Amorphous Feso M7 8 12 CUI (M: Ti, Ta, Nb. Mo) alloys also reveal two
step crystall ization behavior. The DSC curves for M = Mo, Nb and Ti alloys
(Fig. 5.2) show two or three well distinguished exothermic peaks (Miglierini
et aI., 1999). The first peak located between the temperatures of about 450
and 500'C is attributed to the crystallization process consisting in the
nucleation and growth of a new phase. As confirmed by the XRD this peak
corresponds to the formation of crystalline bcc-Fe grains in the residual
amorphous matrix. As before the second exothermic peak (T X2 ) reflects the
complete transformation of the remaining amorphous phase to a mixture of the
coarse bcc Fe grains and iron borides. The nanocrystalline bcc Fe phase can
174 Michat Kopcewicz

300 400 500 600 700


Temperature ('el

Figure 5.1 The DSC curves for Fe93-x-yZr7BxCuy alloys.

be formed in the temperature range between T x, and T x2 The temperatures


of the onset of crystallization, T x, and T x 2 ' are collected in Table 5. 2
(Kopcewicz et aI., 1999a).
Table 5. 1 The crystallization temperatures determined by DSC for Fe93-x- yZr7 B x CUy alloys
( T x, and T X 2 are the onset temperatures corresponding to the first and second crystallization
stage and T, is the temperature corresponding to the minimum in the first crystallization
peak) .

Alloy Tx,('C) T, ('C)

Fes, Zr7 B'2 563 639

FeS5 Zr 7Bs 547 600

FeS7 Zr7B6 546 583

FeS9Zr7 B, 536 557

Fe79 Zr7 B'2 CU2 533 555 713

FeS3 Zr7 B sCU2 522 543 718

FeS5 Zr7 B6 CU2 518 538 720


Mossbauer Spectroscopy Characterization of Soft Magnetic... 175

M=Nb

600 800
Temperature CC)

Figure 5.2 DSC measurements for the as-quenched FeaoM7BI2CuI (M=Mo,Nb and Ti>
alloys (heating rate: 20 K/min).

Table 5. 2 Crystallization temperatures for Feao M7 B I2 CUI alloys determined from DSC
curves.
First peak Second peak

M Onset temperature Peak position Onset temperature Peak position


T XI CC) CC) T X2 CC) CC)

Ta 357 426 568 584


Mo 415 446 658 679
Nb 441 503 727 769
Ti 444 479 719 744

The DSC curves obtained in numerous studies for various amorphous


precursors suitable for the formation of the nanocrystalline phase are very
similar to those presented above, and always reveal at least two steps in the
crystallization process. The nanocrystalline phase can be formed by annealing
at temperatures between the first and the second peak in DSC curves.
5.3. 2. 2 X-ray Diffraction
The X-ray diffraction (XRD) technique, similar to DSC, is routinely used for
the primary characterization of the structure of amorphous precursors, and for
the detection of the formation of the crystalline phase appearing in the
amorphous matrix due to annealing. The XRD is used first of all for the
identification of the crystalline phases, however, in favorable conditions it is
possible to determine the size of grains. This aspect of XRD studies will be
discussed later (see Section 5. 8) .
As an example, the XRD results obtained for the nanocrystallization of the
amorphous
Fe93-x-yZr7BxCuy(x=6, 8 and 12; y=O, 2) (Kopcewicz et aI., 1996,
1997a) and FesoM7B12CUl (M: Ti, Ta, Nb, Mo) (Miglierini et aI., 1999;
176 Michat Kopcewicz

Kopcewicz et al., 2001) are shown.


The XRD measurements were performed for the FeBl Zr7 B12 and F8]g Zr7 B12 -
CU2 alloys (Kopcewicz et aI., 1996) in the as-quenched state and annealed at
550, 600 or 780C . By annealing the FeZrB alloy at 550 and 600 C the bcc-Fe
phase is formed as revealed by the appearance of distinct diffraction peaks
whose positions agree well with the ex-Fe phase (Fig. 5.3). The Fe3(Zr,B)
phase accompanied the dominating bcc Fe phase. In the FeZrBCu alloy only
the ex-Fe phase was formed (Fig. 5. 3f - g). Although not clear in the XRD
patterns, the amorphous phase remains dominant after the 550 and 600 C
anneals as shown by the Mossbauer results below. The Iinewidths of the (110)
peaks in the 600C patterns were used to estimate an average bcc-Fe grain
size via the Scherrer formula (see Section 5.8).

FeSlZr7BI2
(a) As-quenched (e) As-quenched

550'C 550'C

I
I ,I
I
600'C (g) 600'C
I I
I I
I I
I I
I

40 60 80 100 120 40 60
2en
Figure 5. 3 The X-ray diffraction spectra recorded for Fesl Zr7 8 12 and Fe79 Zr7 8 12 CU2
alloys in the as-quenched state a) and (e)) and after annealing at 550'C (Cb) and
(I)),600'C c) and (g)), and 780'C (Cd) and Ch)) Line positions of bcc Fe are
indicated.

At 780C both alloys crystallized completely. The ex-Fe phase dominates


in the FeZrB alloy. Traces of the Fe3 (Zr, B) phase could be detected
(Fig. 5. 3d). In the case of the FeZrBCu alloy clear diffraction peaks
corresponding to Cu appeared in addition to the ex-Fe pattern (Fig. 5.3h).
Similar results were obtained for the remaining Feg3-x- yZr7 B x CUy (x = 6,
8 and 12; y = 0, 2) alloys. In all cases the bcc-Fe was identified as the
Mossbauer Spectroscopy Characterization of Soft Magnetic... 177

nanocrystalline phase (Kopcewicz et al., 1997a). The grain size was


determined for all sample compositions (see Section 5.8 for details) .
Also in the case of FeBoM7B12Cul (M: Ti, Ta, Nb, Mo) alloys, the XRD
patterns corresponding to the FeBo M07B 12 CUI and FeBo Ti 7B 12 CUI alloys
(Fig. 5. 4a,b) consist of broad and diffuse intensity distributions attributed to
the retained amorphous phase and some well-defined but broadened Bragg
peaks assigned to the emerging crystalline grains (Miglierini et aI., 1999).
The as-quenched ribbons are amorphous without any detectable crystalline
phases except for the Ti-containing alloy for which a small peak at about 65
appears (Fig. 5. 4b). Subsequent thermal treatments performed at various
annealing temperatures resulted in the appearance of a crystalline phase
identified as bee-Fe.

580 'C/h 620 'C/h

520'C/h

As-quenched
I I

40 60 80 100 120 40 60 80 100 120


28(") 28(')
(a) (b)

Figure 5. 4 XRD patterns of FeBD M7 8'2 Cu, ribbons: (a) M = Mo, and (b) Ti as-
quenched and annealed for 1 h at the indicated temperatures.

The crystalline phase(s) can be characterized by a structure refinement of


X-ray diffractograms using common methods whereas the average grain size
may be estimated using the Scherrer equation. However, precise quantitative
estimation of the relative fractions of various phases remains difficult
particularly in the case of low volume fraction of the crystalline phase. The
major problem arises from absorption effects, in addition to residual stresses,
texture effects, and instrumental effects.
The XRD technique, similar to DSC, being the standard characterization
method, was commonly used in a majority of investigations of nanocrystalline
alloys for the detection of the onset of nanocrystallization of the amorphous
precursor and for the identification of the chemical form of the nanocrystalline
178 Michal Kopcewicz

phase both in the ease of FINEMET and NANOPERM-type alloys. In almost all
XRD studies of FINEMET-type alloys the bee-Fe ( Si) was identified as the
nanoerystalline phase and in the case of NANOPERM-type alloys the
nanocrystalline phase was bee-Fe.

5. 4 Mossbauer Study of the Structure and Magnetism


of FINEMET Alloys

In 1988 Yoshizawa et al., showed for the first time that annealing the
amorphous Fen 5CUI Nb 3Si 13 5B9 alloy at temperatures exceeding the
temperature of the first step of crystall ization causes the formation of Fe-Si
ultra fine grains with a typical size of about 10- 15 nm. Depending on the Si
content in the Fe-Si grains, which is related to the initial FeCuNbSiB alloy
composition and anneal ing conditions (temperature, time), the Fe-Si phase
reveals the bee structure that for better developed nanocrystallization process
tends to the Fe3Si phase with 00 3 structure. The crystalline nanograins are
embedded in the retained amorphous phase whose composition is affected by
the Fe and Si deficiency. The Mossbauer spectra reflect very well the
development of the crystall ization process. The Mossbauer spectrum of the
starting amorphous FeCuNbSiB alloy, consisting of the broadened sextet
characteristic for the amorphous structure, becomes very complex when the
00 3 structure is formed. It consists of the multiple narrow lines superimposed
on the broad magnetic component. The spectral components with narrow lines
correspond to the structurally different crystallographic sites of crystalline
phases, whereas the broad magnetic component is ascribed to the residual
amorphous phase. The fitting of such spectra is usually difficult because
various spectral components strongly overlap and the hyperfine structure is
poorly resolved. Usually the fitting is performed in terms of the model whieh
associates 4 sextets with narrow lines, with hyperfine fields ranging from about
32 T to about 19 T, to various Fe positions in the 003 structure, and one or
two broad sextets, with hyperfine fields of about 20 and 28 T, corresponding
to the retained amorphous matrix and grain boundary regions. In some cases
the iron borides (Fe23BS or Fe2B) are identified in the crystallized alloy,
expanding further the number of spectral components. The crystallization
process of the amorphous FeCuNbSiB alloys of different compositions has been
extensively studied by the Mossbauer spectroscopy.
Hampel et al. (1992) and Pundt et al. (1992) studied the
nanocrystallization process of the Fen5Cu, Nb 3Si 135 Bg alloy by using the X-ray
diffraction (XRD) and Mossbauer spectroscopy. Changes in the structure of
the amorphous alloy were investigated after annealing for 1 hour at
temperatures ranging from 450 to 800C. It was found that between 520 and
Mossbauer Spectroscopy Characterization of Soft Magnetic... 179

550 C , nanocrystalline Feso Sizo grains with the 00 3 structure were formed
(Hampel et aI., 1992). Grains of about 10 nm in diameter were embedded in
the retained amorphous phase. Above 650C the grains grew up and the
amorphous phase crystallized completely. The iron borate microcrystalline
phases were formed. Kinetics of the amorphous-to-nanocrystalline phase
transformation were investigated by the XRO. The study of the development of
the nanocrystalline phase with time (ranging from 2 min to 1 h) at a fixed
temperature of 520C revealed that most of the nanograins are formed in the
first 10 to 20 min, and already after 5 min. anneal ing the size of the grains
remains almost constant (:::::::;10 nm in diameter). Kinetics of crystallization of
the same alloy were studied also by the Mossbauer effect (Pundt et al.,
1992). A similar study was performed by Cserei et al. (1994), who
determined the activation energy for the crystallization of 143 kJ/mol.
Crystallization of the amorphous Fen5 CUI Nb 3 Si l65 B6 alloy was studied by
Mossbauer and XRO techniques by Jiang et al. (1991). Formation of the Fe-Si
nanograins with the 00 3 structure was also observed. The Mossbauer spectra
were fitted using similar assumptions with 5 sextets, 4 of which were assigned
to Fe in various positions in the 003 structure, and the fifth, with a small
hyperfine field of 6.5 T, to Fe atoms in the highly distorted environment
remaining between the nanocrystals. A small value of the hyperfine field in the
grain boundary suggested that these regions are enhanced with B, Cu and Nb
atoms. The difference in the magnetization orientation at the sample surface
(in-plane al ignment) and in the bulk (normal to the plane of the sample) was
detected by conversion electron and transmission Mossbauer measurements.
Crystallization behavior of amorphous Fen5 CUI Nb 3 Si 135 B9 alloy was
studied by complementary techniques including the Mossbauer effect by
Rixecker et al. (1992), who also formed a favorable nanocrystalline Fe(Si)
structure mixed with a remaining amorphous fraction in a sample annealed at
550C for 1 h.
Zhou et al. (1993), studied the nanocrystallization of the Fen 5CUI Nbr
Si 135 B9 alloy, and similarly to the above mentioned investigations, observed
the formation of nanograins of Fe (Si) with 00 3 structure. It was found that
when the amorphous residual phase becomes paramagnetic (the estimated
Curie temperature for this phase was about 600 K), the coercivity of the alloy
dramatically increases. It follows that in addition to the refinement of the size
of nanograins, the coupling between the grains through the amorphous matrix
is important for the achievement of the excellent magnetic softness. The
exchange coupling between the grains was studied for the two-phase
Fe695Cuo.5Cr4 V5Si 13 Bs nanocrystalline alloy (Yang et aI., 1999). A
phenomenological model of coupling was applied to estimate the coupling
intensity for samples annealed at different temperatures. The strongest
coupling was found for the alloy annealed at 510 - 540C explaining the best
soft magnetic properties observed at 540 C .
Structural and magnetic changes in Fen 5CUI Nb 3 Sizz- x B x (x = 6 and 9)
180 Michat Kopcewicz

amorphous alloys during the crystallization process were studied in detail using
the XRD and M6ssbauer spectroscopy by Gorria et al. (1996). Authors
concluded that the amount of the nanocrystalline FeSi phase is different for
each composition but it does not change between the end of the first
crystallization step and the full crystallization of the samples. The Si content in
the nanocrystalline phase remains almost unchanged and is close to the total
content of Si in the original alloy. The Curie temperature for the retained
amorphous phase was unexpectedly high due to compositional inhomogeneities
in the amorphous matrix and the exchange field penetration from the FeSi
grains. However, the Curie temperature of the Fe-Si nanophase was found to
be lower than expected; that was attributed to the defects in the FeSi
nanocrystals and the influence of interfacial regions between amorphous and
nanocrystall ine phases.
Pradell et al. (1995) have used the M6ssbauer spectroscopy to study the
nanocrystallization process in the Fe73 5CU, Nb3Si175B5 alloy during isothermal
anneal ing at 490 'c and observed two distinct stages: CD consisted in changes
of the short range order in the amorphous phase, and CV nanocrystallization of
00 3 Fe-Si phase. The first stage lasts only few minutes after isothermal
annealing and is accompanied by an increase of the hyperfine field in the
amorphous phase. The hyperfine field decreases during the second stage. The
Si-content in the nanocrystals decreases with annealing time. The grain size
distribution strongly suggests that nanocrystallization is driven by nucleation
and growth. During annealing the magnetic moments align preferentially in the
plane of the sample, especially in the first stage of crystallization.
Thermodynamic and kinetic factors controlling the formation of the
nanocrystalline Fe735CulNb3Si,75B5 alloy was studied by XRD, M6ssbauer
spectroscopy and transmission electron microscopy by Clavaguera et al.
( 1995). Their results show that a 00 3 structure develops directly from the
amorphous matrix. The first stage of the crystallization process under both
isothermal and continuous heating regimes is controlled by nucleation and
growth. Further nanocrystallization is controlled by diffusion limited growth.
Kinetics of nanocrystallization of amorphous Fe725-xCulNbuSilO+x+yB12-y
alloys were studied also by Miglierini et al. (1994a), who have shown that
compositional changes in the FeCuNbSiB system affect the crystallization
kinetics. Crystallization results in a segregation of the alloy constituents and
affects the short range order of the original amorphous structure.
Short-time high-temperature annealing of the Fen 5CUI M03Si,3 5Bg alloy
was compared with the conventional Nb-containing FINEMET alloy (Girchardt
et aI., 1997).
Influence of neutron irradiation on the short range order of amorphous and
nanocrystalline Fen5Cu, Nb 3Si 135 Bg alloy was studied by the M6ssbauer
spectroscopy (Miglierini et aI., 1994b) and DSC, XRD and magnetic
measurements (Skorvanek and Gerl ing, 1992). The M6ssbauer investigations
revealed a redistribution of the alloy constituents after neutron irradiation.
Mossbauer Spectroscopy Characterization of Soft Magnetic... 181

Radiation damage introduced macroscopic stress centers, which together with


positive magnetostriction of the amorphous phase forced the magnetization to
turn out of the ribbon plane. The structure modifications of amorphous
Fe30Ni36Cr12M02Si5B16 alloy induced by neutron irradiation have also been
studied by M6ssbauer spectroscopy (Skorvanek and Miglierini, 1991).
Temperature dependence of magnetization was studied in a wide
temperature range for the as-quenched amorphous and nanocrystalline
Fe735 CUI Nb 3 Si 165 B6 alloy (Gupta et aI., 1995). It was found that at low
temperature the temperature dependence of magnetization arises from
excitations of long-wavelength spin waves for the amorphous alloy but does not
do so for the nanocrystalline alloy. For short annealing times (2 min.), i. e.,
for small volume fraction of nanocrystalline grains, a superparamagnetic
behavior was observed for the grains above the Curie temperature of the
amorphous matrix.
The M6ssbauer spectra of the Fe66 Cra CUI Nb 3 Si 13 B9 nanocrystalline alloy
were measured as a function of temperature for various annealing times
(Randrianantoandro et al., 1995). The results suggest that a segregation of
atoms takes place during the crystallization process leading to the enrichment
in Cr of the residual amorphous matrix and to the formation of Fe-14 % Si
nanograins. Magnetic interaction in the same alloy has been studied at various
temperatures by means of the M6ssbauer spectroscopy (Ranchanantoandro et
al., 1997). The suppression of the interphase exchange interactions at the
ferromagnetic to paramagnetic phase transition of the amorphous matrix and
superparamagnetic fluctuations were observed.
In the papers discussed above the analysis of the M6ssbauer spectra was
focused on the nanocrystalline phases. A somewhat different approach was
proposed by Miglierini (1994), who concentrated in his analysis on the
changes in the hyperfine field distributions, calculated from the M6ssbauer
spectra of amorphous and nanocrystalline FeCuNbSiB alloys, which yield
information on the structure of the residual amorphous phase. The spectra
were fitted by a combination of the component related to the distribution of
hyperfine fields and discrete spectral components with narrow lines originating
from various Fe positions in the crystalline Fe(Si) phase. The hyperfine field
distributions, P (H), were usually extracted from the M6ssbauer spectra using
a histogram method (Hesse and Rubartsch, 1974; LeCaer and Dubois, 1979).
The dependence of the hyperfine parameters on the time of annealing the
Fe735-xCuINb3+xSi135B9(X=Q, 1.5) alloys and a different behavior of the
intergranular component in both samples were studied via three-dimensional
(3D) projections of the hyperfine field distributions which allow easy
visualization of the changes in the P (H) vs. annealing time. The P(H)
distributions clearly show a bimodal nature due to two distinct Fe atom
environments. The high-field part of P (H) can be associated with Fe atoms
which have primarily Si and B as nearest neighbors, whereas the low-field part
is ascribed to Fe atoms surrounded by Cu, Nb and B atoms. The relative
182 Michat Kopcewicz

contribution of the high and low-field parts to the total P (H) distribution
changes with progressing crystallization. With the increase of annealing time
and segregation of Fe, Si and B atoms from the original amorphous precursor
reduce the number of Fe atoms with large hyperfine fields. The low-field
component dominates for long annealing times. The importance of the
intergranular component is also noticed in this paper, but, as in all
investigations discussed here, a relevant detailed information cannot be
obtained from the Mossbauer spectra because of their complexity related to
many Fe positions in the 003 structure and in the retained amorphous matrix.
This problem will be discussed in detail in the next section, devoted to
NANOPERM-type alloys.
Borrego et al. (2000) studied several series of nanocrystalline
FeSiBCuNbX alloys (X = Zr, Nb, Mo, V). By applying a refined fitting
procedure (which basically resembles the models discussed above) the
hyperfine field distributions were compared for various X atoms. A similar
behavior was found for different alloy compositions for a given crystalline
fraction indicating that the kinetics and diffusion mechanisms are independent
of X element. The hyperfine field distributions show a bimodal shape when a
crystall ine fraction exceeds 25 %, suggesting an inhomogeneous residual
amorphous phase. The evolution of magnetic texture with the crystalline
fraction was followed.
Conventional Mossbauer studies discussed above allow the identification
and estimation of the relative abundance of phases formed due to annealing of
the amorphous precursor such as the nanocrystalline bee-Fe ( Si) phase,
retained amorphous matrix and sometimes the intergranular regions. They do
not provide information about the grain size, which must be obtained from
other methods, e. g., transmission electron microscopy and X-ray diffraction.
Also information regarding magnetic properties, such as the magnetic
anisotropy fields, coercivity and magnetostriction, are not available from
conventional Mossbauer measurements. Therefore, the nonconventional
rf-Mossbauer technique, in which the rf collapse and rf sidebands effects are
employed, was used to distinguish magnetically soft nanocrystalline and
amorphous phases from magnetically harder microcrystalline phases formed
due to annealing, and to obtain information about magnetic anisotropy fields
and magnetostriction of each phase present in the multicomponent
nanocrystall ine alloys.
The first study of the nanocrystall ine alloy, in which the rf-Mossbauer
technique was applied, was performed for the Fe73 5 CUI Nb 3 Sin 5 B9
(FINEMET) alloy (Kopcewicz et al., 1994a). Let us discuss these results in
more detail.
Anneal ing of the amorphous Fen 5 CUI Nb 3 Si 13 5 B9 alloy induced substantial
changes in the alloy's microstructure which are clearly seen by the
conventional Mossbauer technique. The spectra were measured for the original
as-quenched sample, for which a typical spectrum of an amorphous alloy with
Mossbauer Spectroscopy Characterization of Soft Magnetic... 183

the average hyperfine field of 21 . 5 T was recorded (F ig. 5. 5a) , as well as for
the annealed samples (Fig. 5. 5b-f) which reveal sharp lines belonging to the
4 or 5 sextets. Hyperfine parameters allowed us to identify the nanocrystalline
Fe-Si phase with D0 3 structure (Fig. 5. 5b - d) and Fe2 8, Fe38 as well as
microcrystalline Fe-Si phases in the spectra shown in Fig. 5. 5e, 1. The
spectra were fitted in the same way as discussed above (Pundt et al., 1992,
Hampel et ai., 1992). Annealing at 520 - 570 'c causes the formation of
nanocrystall ine ex-Fe (Si). The remaining amorphous phase reveals a smaller
magnetic hyperfine field of about 19 T (Fig. 5.5b, c, d). Annealing at
temperatures higher than 600 'c leads to the conventional crystallization of the
amorphous Fe735CulNb3Si1358g alloy, and as a result the crystalline bcc-
Fe(Si) and tetragonal Fe38 and Fe28 phases appear (Fig. 5. 5e,f).
The spectra of the as-quenched and annealed samples were measured
also during exposure to the rf field of 20 Oe at 60. 5 MHz. In the as-quenched
amorphous alloy the rf field induced the complete rf collapse of the magnetic
hyperfine structure, as evidenced by the central doublet in the spectrum
(Fig. 5. 5g). Strong sidebands, accompanying the doublet, show that the
alloy studied is highly magnetostrictive. The spectra recorded during the rf
exposure for the annealed samples differ dramatically from that observed in
Fig. 5. 5g for the as-quenched alloy. The sample annealed at 520 'c consists
of nanocrystalline grains embedded in the remaining amorphous phase. This
alloy exhibits soft magnetic properties, which cause the preferred orientation
of magnetization to be parallel to the absorber plane yielding the line intensity
ratio of 3 : 4: 1 (Fig. 5.5b). The spectrum (Fig. 5.5h) recorded for this
sample during rf exposure reveals the rf collapsed doublet, related to the
remaining amorphous phase, and rf sidebands. which are, however, weaker
than those for the as-quenched sample. This indicates that the
magnetostriction of this sample becomes smaller. Additionally we observe a
non-collapsed spectral component with magnetically split lines. The
magnetically split part is most probably related to the nanocrystalline phase.
The non-collapsed spectral component indicates the existence of a local
magnetic anisotropy. In Fig. 5. 5c one can see that the content of the
remaining amorphous phase has decreased. Measurement of this sample in the
rf field shows a further decrease of the sidebands' intensity (Fig. 5.5j) and a
more complete collapse of the spectrum representing the nanocrystalline
phase. This trend continues. as seen in Fig. 5. 5d, j. The sidebands disappear
completely and the triangular, partly collapsed spectral component narrows
further and its spectral contribution decreases. These three examples (Fig. 5.
5h, i. i) show that the rf field influences the remaining amorphous and
nanocrystalline parts in different ways. Thanks to this phenomenon we are able
to perform a selective M6ssbauer investigation for the different phases in
question. Additionally, this method allows us to obtain information about the
magnetostriction and the local magnetic anisotropy simultaneously.
184 Michat Kopcewicz

No rf field In rf field
1.00
1.00
""'";r--..::!"""'\Vr-
'.
0.95
...
.. ,
(g)
0.95
1.00
1.00 ,~

'.
0.95 (h)
0.95
1.00

\T
1.00

'o" '"
:.. :
'iii .~ 0.95
'E'" 'E'" (i) tt
(/) 0.95 '"
~ 1.00
~ 1.00
'";>
.~
'"
;>
.~
Q) Q) 0.95
~ e<:
0.95 (j) t,;
1.00 1.00

0.95
1.00 0.95
1.00

(t)
0.95 '----_-L--*----'--_---'-_----' 0.95 L - _ - - ' - - _ - - - - - - l_ _--'--_---'
-10 -5 o 5 10 -10 -5 o 5 10
Velocity (mm/s) Velocity (mm/s)

Figure 5.5 The M6ssbauer spectra recorded for the as-quenched (a) and annealed (b) -
(f) Fen5 CUI Nb 3 Si ,3 589 alloy prior to the rf exposure, and for the same samples exposed
to an rf field of 20 Oe at 60.5 MHz (9) - (I).

The situation changes completely for the samples annealed at T>600 'c .
Figure 5. 5e shows the spectrum of an alloy consisting of microcrystals. Here
the magnetization is not aligned in the absorber plane. The line intensities ratio
is about 3 : 2 : 1 which indicates that the alloy has lost its soft magnetic
properties. The increased magnetic anisotropy related to the formation of
Fe3 B and Fez B suppresses the rf collapse of hyperfine fields. The central part
of the spectrum recorded during the rf exposure (Fig. 5. 5k) comes mainly
from the overlapping magnetically split subspectra of Fez B, Fe3 Band Fe-Si
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 185

phases, and contains only marginal rf collapsed component related to the


Fen5 CUI Nb 3Si 135 B9 retained amorphous fraction which contributes to about
20 % of the total spectral area observed in Fig. 5. 5e. The spectrum recorded
for this sample (Fig. 5.5k) is very similar to that recorded for the sample
annealed for 4h at 750'C (Fig. 5.51) in which the nanocrystalline phase was
not formed and the crystal Iization products the Fe2 Band Fe3 B phases
appeared in addition to the Fe-Si microcrystals of 003 structure. The retained
amorphous component contributing to less than 20 % of the total spectral area
is still observed (Fig. 5.50.
The rf exposure does not affect the microstructure of the
Fen5 CUI Nb 3Si,35 B9 alloy. The spectra recorded for all the as-quenched and
annealed samples before (Fig. 5. 5a - f) and after rf exposure were
identical.
The results show that the rf collapse and sideband effects are very
sensitive to the microstructure of the alloy and allow us to follow the changes of
the structure due to anneal ing. The onset of the nanocrystall ine phase affects
the rf collapsed spectra by the formation of a partially collapsed spectral
component of a triangular shape and the decrease, and finally the
disappearance of the rf sidebands due to the decrease of the magnetostriction.
Conventional crystallization of the amorphous Fe73 5 CUI Nb 3Si 135 B9 alloy at
T>600'C leads to the formation of crystalline phases with large magnetic
anisotropy which prevents the magnetization reversal and suppresses the rf
collapse. Hence, it is possible to distinguish clearly, by studying the rf
collapse and sidebands effects, between crystallization leading to either nano-
or microcrystalline phases, whose composition can be identified by the
Mossbauer technique. Various phases present in the annealed alloy lead to
different spectral components: the completely collapsed doublet corresponding
to the retained amorphous phase, the partially collapsed, triangular
component characteristic of the nanocrystall ine phase, and the non-collapsed,
magnetically split patterns corresponding to the crystalline Fe3B and Fe2B
phases and microcrystalline Fe-Si.
Further rf-Mossbauer investigations of FINEMET alloy (Graf et ai., 1995a,
1995b, 1999) revealed the changes of the magnetostriction vs. annealing
temperature. By studying the relative sideband intensities in the rf-collapsed
spectra of Fen5Cu,Nb3Si135B9 alloy annealed at temperature range of
400 - 600 'c, one can follow the changes of magnetostriction since the
normalized average sideband area is proportional to the magnetostriction of
the sample. The sidebands' area and hence the magnetostriction rapidly
decrease already at the onset of crystallization and vanish as the
nanocrystalline phase is developed. These data agree well with the
macroscopic measurements of magnetostriction of this alloy (MOiler et al.,
1992). The structural changes induced by annealing in FINEMET alloy, studied
by the rf-Mossbauer technique were discussed by Graf et al. (1995b).
Influence of various heat treatments on the formation of the soft magnetic
186 Michat Kopcewicz

nanostructure in the amorphous Fe73 5CUI Nb3Sil35Bg alloy was studied by the
experiments performed in which the short-time anneal ing at elevated
temperatures was compared with 1 hour annealing at 550 'c , the temperature
commonly used for the formation of the nanocrystalline FINEMET alloy for
technical applications. It was found (Graf et al., 1996) that it is possible to
obtain a good nanocrystalline phase in a simpler, less time consuming process
in which a short-time annealing (15 - 120 s) is performed at 550 - 650 'c
i. e. , at temperatures at which 1 h heat treatment leads to the deterioration of
soft magnetic properties. The Fen5 CUI Nb 3Si l35 Bg samples annealed for 15-
300 s at T a = 550 and 650 'c were characterized by the rf-Mossbauer and
electron microscopy techniques. The criterion for the" good soft magnetic
properties", i. e., comparable to those achieved in the conventional heat
treatment (1 h at 550 'c ), was that the rf sidebands vanish in the rf-Mossbauer
spectra and the rf-collapse of the magnetic hyperfine structure is complete.
The rf-Mossbauer spectra measured for the short-time annealed samples were
compared with the corresponding spectra obtained for the samples annealed
for 1 hat 550-650 'c (Graf et al., 1996). Investigation of the annealing of
the Fen5CuINb3Sil3 5B9 alloy at 550 'c as a function of time from 30-120 s
shows that even 60 s of annealing is sufficient for the formation of the Fe(Sj)
phase (F ig. 5. 6b) which has local magnetic anisotropy small enough that full rf
collapse of the magnetic hyperfine structure can occur (Fig. 5.6b'). Shorter
anneal ing (Fig. 5. 6a') leads to the formation of Fe ( Si) with substantial
anisotropy resulting in the partly collapsed spectral component in Fig. 5. 6a'
similar to that observed for 1 h annealing at about 520 'C. However,
magnetostriction is greatly reduced as is evident by the disappearance of the rf
sidebands in Fig. 5. 6a'. Annealing at 550 'c for 60 - 120 s produces a very
soft nanocrystall ine phase as shown by the rf-Mossbauer spectra in Fig.
5. 6b' - d' which cons is of a single line rf collapsed pattern; the rf sidebands
vanished completely. Only the rf-Mossbauer experiment is sensitive enough to
reveal the changes in magnetic properties (anisotropy, magnetostriction)
related to the structure of nanoscale gra ins. The conventional Mossbauer
spectra, recorded in the absence of the rf field, do not differ much from each
other (Fig. 5. 6a - d). A nanocrystalline phase can also be formed with further
shortening of the annealing time at elevated temperature (15 sat 650 'C). The
finding that the nanocrystall ine bee-Fe (Sj) phase with excellent soft magnetic
properties and vanishing magnetostriction can be formed in the
Fe735CuINb3Sil35Bg alloy via short time annealing at 550-650 'c is important
from the point of view of technical applications of such materials. Long, 1 h
annealing seems to be unnecessary, thus making the process of the formation
of nanocrystals simpler and less time-consuming.
The unique rf-Mossbauer technique is very sensitive to the changes in the
microstructure of the alloy studied and allows the determination of the magnetic
Mossbauer Spectroscopy Characterization of Soft Magnetic... 187

No rf field In rf field
1.00
,.. ~,
1
1.00 ~
7
: ""'"

"'"iI'. .....
!-. ,
,...
: 'If ,
...,.;
1'1 .
(a)
"!~
. ~;..
. 30 s
0.95
(a') "I
0.95
1.00 nil
1.00
~F
..
.
t: t:
o
'12 (b) 60 s .~ 0.95 (b')
.~ 0.95
g 1.00 ..., 'E
~
if>
1.00 .......c" ....'"
l.!

~f1, ~ '">
t.~
:It:
: ~
.,.
~

.;'.
....
(c)
-,.,Jlo .,:
.:- 90 s
0::: 0.95
(c')
,
0.95
- . . ... ~ .... "
1.00
~ r- 1.00
l.!
W'~ :.; .;-..,., !'tl\
". ...
.:':~ ...:
:)~ ...
(d) ; 120 s 0.95 (d')
,
-8 -4 o 4 8
Velocity (mm/s) Velocity (mm/s)

Figure 5.6 M6ssbauer spectra recorded for Fen5 Cu, Nb 3 Si 135 8 9 for samples annealed at
550C as a function of time from 30 s to 120 s without (a) - (d) and with (a') - (d') the
rf field.

properties (anisotropy, magnetostriction) of the phases formed due to


annealing, thus permitting the distinction between the magnetically soft
amorphous and nanocrystall ine phases, and magnetically harder
microcrystalline grains. This technique is clearly superior to the conventional
Mossbauer studies, which only allow the identification of phases. However,
information concerning the absolute sizes of the grains cannot be directly
extracted from the Mossbauer studies. The transmission electron microscopy
(TEM) measurements are indispensable for the grain size determination.
However, the TEM technique does not allow the identification of the
composition of the phases formed and of their magnetic properties. The
systematic study with the use of complementary methods are necessary for
better understanding of the structural and magnetic properties of the
nanocrystalline phases appearing in the course of annealig the amorphous
precursor.
188 Michat Kopcewicz

5.5 Mossbauer Study of the Structure and Magnetism of


NANOPERM Alloys

In 1990 a new type of nanocrystalline alloy was developed (Suzuki et al.,


1990). It is a ternary Fe-M-B alloy (M: Zr, Hf, Nb, Ta, ... ) with nanoscale
grains of mostly single phase of bcc-Fe. The bcc-Fe nanocrystalline phase has
been prepared by annealing the amorphous FeZrB, FeHfB and various FeZr-M-B
(M-transition metal) alloys (Suzuki et al., 1991a, 1991b, 1993; Makino
et al., 1995, 1997a, 1997b). The nanophase is formed by annealing the
relevant amorphous precursor, which reveals two-stage crystallization
behavior As in the case of FINEMET-type alloys, a soft magnetic nanophase
can be formed during primary crystallization. Soft magnetic properties
deteriorate dramatically after complete crystallization is achieved, resulting in
the increase of the grain size and formation of magnetically harder phases, e.
g., iron borates. The nanocrystalline alloys consisting of the bcc-Fe grains
embedded in the retained amorphous phase, exhibit high saturation
magnetization (-1.7 T) combined with very low anisotropy and coercive
fields and almost vanishing magnetostriction. The magnetic properties of such
alloys are better than those of FINEMET alloys; in particular the saturation
magnetization is significantly higher. The 1 % - 2 % Cu content in Fe-M-B
alloys causes, as in the case of FINEMET, the decrease of the crystall ization
temperature of the amorphous precursor and increase of the nucleation rate of
the crystalline phase. Boron enhances thermal stability of the amorphous
phase (Suzuki et al., 1993) and affects the homogeneity of the bcc
precipitates (K im et aI., 1994). Since the first development of ternary
nanocrystall ine alloys, many new compositions were prepared and studied.
For a recent review of the structural and magnetic characterization with various
experimental techniques of this class of nanocrystall ine alloys, called
NANOPERM, see, e. g., Refs. (Makino et al., 1997a, 1999; McHenry et aI.,
1999) .
The relation between saturation magnetization (B s ) and effective
permeability at 1 kHz for nanocrystalline Fe-M-B based alloys, FINEMET
alloys, conventional amorphous alloys and crystalline soft magnetic materials
are shown in Fig. 5.7 (Makino et al., 1997a). It is expected that NANOPERM
alloys will find wide practical applications, especially in magnetic devices
such as power transformers, choke coils, pulse transformers, flux gate
magnetic detectors, magnetic heads, sensors, magnetic shielding, etc.
Figure 5. 8 summarizes expected appl ication fields for NANOPERM alloys
together with the magnetic characteristics required for their applications
(Makino et aI., 1999).
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 189

Co-Based amorpholls alloys

2 >< 105

1 X 10 5
Nanocrystalline
Fe-M-B based alloys
5x 10'

I X 10'

5>< 103
Fe-Based
Amorphous alloys
2x !O3
Mn-Zn Ferrite
1x 103 LI -'----_ _____' -'-- ~_ _____'_ _

o 0.5 1.0 1.5 2.0 2.5


Bs(T)

Figure 5.7 Relation between B, and Ile at 1 kHz for the nanocrystalline Fe-M-B based
alloys, nanocrystalline Fe-Si-B-Nb-Cu alloys and conventional soft magnetic materials
CMakinoetal., 1997a).

The structure, stability and magnetic properties of NANOPERM alloys


were characterized by similar experimental techniques as in the case of
FINEMET alloys. The crystallization temperatures in amorphous precursors
were usually established using the DSC measurements; the identification of
crystalline phases formed as a result of annealing the amorphous precursor was
done using the XRD and M6ssbauer spectroscopy; the size of nanograins was
commonly determined by transmission electron microscopy (TEM) and X-ray
techniques (XRD). Sometimes and small angle X-ray scattering (SAXS)
technique was applied. The M6ssbauer spectroscopy was widely applied by
numerous researchers to study structure and magnetism of NANOPERM alloys.
The most popular systems which recently attracted considerable attention were
Fe-M-B(Cu) (M: Zr, Nb, Mo, Ti, Ta, ... ) alloys of various compositions.
Unlike in Si-containing FINEMET-type alloys, in which the nanocrystalline
Fe(Si) phase yields many-component complex M6ssbauer spectra difficult to
fit and interpret, the spectra of NANOPERM-type alloys are much simpler.
They consist of only one sextet related to the nanocrystalline phase (usually
bcc-Fe) superimposed on a broadened pattern originating from the retained
amorphous matrix and possible interfacial regions. In simple cases the broadened six-
190 Michal Kopcewicz

Power transformers
Pole transformers
Power transformers for
switched power suppl ies

Data communication
interface components
Pulse transformers

EMI prevention components


Common mode choke coils

Magnetic heads

Sensors
Current transformers
Magnetic direction sensors

Magnetic shielding

Reactors
Magnetic saturable choke coils
Magnetic switching cores

Figure 5.8 Magnetic characterizations and application fields for the Fe-M-B based alloys,
NANOPERM (Makino et aI., 1997a).

line pattern was fitted with the hyperfine field distribution method. In early
studies, one P (H) distribution was found sufficient for achieving a reasonable
fit, but since in many cases the P (H) distributions revealed well separated
low and high-field parts, the fit with two independent P (H) distributions were
made. One P (H), with smaller average hyperfine fields, was usually
attributed to the retained amorphous matrix, whereas the second P (H)
distribution, with larger hyperfine fields, was related to the interfacial regions
between the crystalline nanograins and the amorphous matrix. In all studies the
formation of the nanocrystalline phase was identified from the appearance in
the Mossbauer spectrum of the sextet with narrow Lorentzian lines and
characteristic hyperfine field of about 33 T and isomer shift {) = O. 00 mm/s,
corresponding to the bee-Fe phase. The relative spectral area of this sextet
was used for evaluation of the volume fraction of the nanocrystalline phase in
the composite alloy. Usually, such estimates contain a systematic error
because the possible differences in the Debye-Waller factors for various
phases (nanocrystalline, amor'phous, interfacial) were not taken into account.
The Mossbauer results obtained for various NANOPERM-type
nanocrystalline alloys will be reviewed in two sections: CD in which the
conventional Mossbauer studies (i. e., the measurements in transmission
geometry without external fields, Section 5. 5. 1) and (2) in which the
unconventional rf-Mossbauer measurements (Section 5. 5. 2) are presented.
The surface phenomena (surface crystallization, surface spin texture) are
Mossbauer Spectroscopy Characterization of Soft Magnetic... 191

discussed in Section 6.

5.5. 1 Conventional Mossbauer Studies


Soon after the development of NANOPERM alloys (Suzuki at a!., 1991 a,
1991b; Makino et a!., 1995) the Mossbauer spectroscopy was applied for the
study of their properties. The nanocrystalline Fe86 Zr7 B6CUI alloy was studied
by Gorria et a!. (1993) and Orue et a!. (1994). The nanocrystalline phase was
obtained by annealing the amorphous precursor at temperatures 650 - 870 K.
The Mossbauer spectra were recorded at 77 K, room temperature and at 360 K
(above the Curie temperature, T c' of the amorphous phase). It was found
from the spectrum at the temperature exceeding T c that in the as-quenched
amorphous alloy two inequivalent Fe sites exist. Below T c' a bimodal
hyperfine field distribution indicated two Fe environments. A similar feature
was observed also in the remaining amorphous phase in the nanocrystalline
alloy. In the first step of crystallization the ex-Fe precipitates and its content
increases with increasing annealing temperature. However, the conclusion that
the overall composition of the sample does not change noticeably is rather
strange in view of the marked increase of the relative fraction of almost pure
ex-Fe, leading to the deficiency of Fe in the residual amorphous phase. At high
annealing temperatures (790 and 870 K) the P (H) distributions change,
suggesting the formation of Fe-B crystalline phases, which, however, has not
been confirmed by XRD measurements.
Navarro et a!. (1995), investigated local structural changes during the
crystallization process of the amorphous Fe88Zr7B4Cul alloy. This alloy is
paramagnetic at room temperature, but the precipitation of ex-Fe, occurring
due to annealing at 500 - 650'C leads to the appearance of the magnetic
sextet in the Mossbauer spectra. The remaining amorphous matrix evolves to a
ferromagnetic state that is related to the increase of the Curie point for the
amorphous phase with lower Fe content. The average hyperfine field of the
residual amorphous matrix increases to about 15 - 22 T with increasing
annealing temperature, T A Annealing at temperatures exceeding 700'C
causes complete crystallization of the amorphous phase and new crystalline
phases are found. The amount of ex-Fe phase increases from less than 60 % for
T A = 650'C to 73 % for T A = 705 C. The Mossbauer results indicated that the
gradual improvement in the magnetic properties of the nanocrystalline alloy is
related to the increase of the volume fraction of ex-Fe phase and the resulting
changes of composition of the retained amorphous matrix (from Fes8 Zr7 B4CUI
for the as-quenched alloy to Fe66Zr19.8Bl1.3CU28 after annealing at 650 C).
Nanocrystallization of the amorphous Fe81 Zr7B12 and Fe79Zr7BI2Cu2
(Kopcewicz et a!., 1995a, 1995b, 1996) alloys revealed also the formation of
the bee-Fe phase at annealing temperatures exceeding 500'C and the increase
of the volume fraction of this phase with increasing annealing temperature. In
192 Michat Kopcewicz

these studies the rf-Mossbauer technique was used (see Section 5. 5b) in
addition to conventional Mossbauer experiments. Then the microstructure and
magnetic properties of a whole group of nanocrystalline
Fe93-x-yZr7BxCuy(x=6, 8 and 12; y=O, 2) alloys was investigated
(Kopcewicz et ai., 1995c, 1997a).
A whole set of FeZrBCu alloys was studied systematically as a function of
alloy composition using various experimental techniques including conventional
and rf-Mossbauer methods. These results will be reviewed in more detail, as
an example of what we can learn about nanocrystalline alloys by using the
Mossbauer spectroscopy (Kopcewicz et ai., 1996, 1997a, 1997b). Let us
discuss first the Mossbauer results obtained for the Fe81 Zr7 B I2 and
Fe79 Zr7 B I2 CU2 alloys (Kopcewicz et al., 1996, 1997a). Both amorphous
alloys crystallize in two steps, as determined by DSC measurements (see
Section 5. 3. 2. 1). Upon anneal ing the crystall ine bcc-Fe phase is formed, as
revealed by XRD measurements, which allow determination of the bcc-Fe
grain size (see Section 5. 8). The conventional Mossbauer measurements
performed at room temperature reveal clearly the changes in the
microstructure of the amorphous FeZrBCu alloy induced by annealing. All
Fe93-x-yZr7BxCuy(x=6, 8 and 12; y=O, 2) alloys were fully amorphous in
the as-quenched state. Annealing at T A = 430 'c does not induce any
detectable crystallization. In all alloys except for that with x = 12 and y = 0,
annealing at T A = 500 'c induces the crystallization (this temperature exceeds
the onset temperature of the first peak in DSC curve, corresponding to the first
stage of crystallization process, Fig. 5. 1) and the bcc-Fe phase appears. The
typical results are shown in Figs. 5. 9 and 5. 10 for the Fe81 Zr7 B I2 and
Fe79 Zr7 B I2 CU2 alloys, respectively. The spectra were fitted either by the
hyperfine field distribution (P(H method (Figs. 5.9a'-c' and 5.10a'-b')
or by combining the P(H) distribution and a subspectrum with a discrete value
of the hyperfine field equal to 32. 95 T and isomer shift {j = 0 mm/s
corresponding to crystalline ex-Fe, (Figs. 5. 9d' - e' and 5. 10c' - e'). The
hyperfine field distributions were extracted from the experimental spectra by
using the constrained Hesse-RObartsch method (Hesse and RObartsch, 1974,
LeCaer and Dubois, 1979). To account for the asymmetry clearly seen in the
magnetically split spectra (Figs. 5. 9a-c and 5. lOa-b), a linear correlation
between the hyperfine field and the isomer shift was assumed. The P ( H)
distributions extracted from the spectra of both alloys in the amorphous state
consist of a single, fairly symmetric bell-like peak (Figs. 5. 9a' - c' and
5. lOa' - b').
When the Fe81 Zr7 B I2 alloy is annealed for 1 h at temperatures below
500 'c , crystallization is not yet started (Fig. 5. 9a - c). However, such an
annealing induced a structural relaxation in the amorphous state in both
samples. The average hyperfine field, calculated from the P ( H) distribution
shown in Fig. 5. 9a', increased from 16.2 T for the as-quenched Fe81Zr7BI2 to
17.7 T after annealing at 500 'c (Fig. 5. 9c'), and the D23 parameter, defined
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 193

0.95

<::
0
.;;;
'2'"
'"<::
~
Q)
> 0.95
.~
0)
0:::

-6 0 6
Velocity (mm/s)

Figure 5.9 The Mossbauer spectra and corresponding P (H) distributions for the Fe81 Zr7 8 12
alloy in the as-quenched state (a, a') and after annealing at 430'C (b, b'), 500'C (c, c'),
550'C (d, d') and 600'C (e, e').

as the intensity ratio of the second line in the sextet with respect to the third
line, which gives information about the average orientation of spins in the
sample, increased from 2. 97 to 3. 40, thus suggesting that after annealing at
500 'c the spins are al igned closer to the plane of the ribbon. The structural
relaxation effect in the Fe79 Zr7 B 12 CU2 alloy was less pronounced
(Fig. 5. lOa' - b').
The first stage of the crystallization strongly affects the microstructure of
the alloys. Annealing the FeS1Zr7B12 alloy at T A ;? 550 'c causes partial
crystallization of the amorphous alloy as a result of which the bee-Fe phase is
formed. This process is clearly observed in the Mbssbauer spectrum by the
appearance of a six-line spectral component with sharp lines corresponding to
194 Michat Kopcewicz

1.00
."
5
.
0.96 (a')

0 ..'
.............
."
1.00

5

.. .............
(b')
0.96
.. ..
'
. .
0
g
.;;;
1.00
V>
.
5

~ ~
........ (c')

'" .... ..
...... ......
<I)
;>
.~ 0.96
~ 0

.
1.00
4 . .
'
"

.... .
(d')

0.96 0 ....
1.00 4
..
..'. .......
(e)
..'
.. '
(e')

0.95
600'C
.. '

--fi 0 6 0 20 40
Velocity (mm/s) H(T)

Figure 5. 10 The Mossbauer spectra and corresponding P ( H) distributions for the


Fe79 Zr7 8 12 CU2 alloy in the as-quenched state (a, a') and after annealing at 430'C (b,
b'), 500'C (c, c'), 550'C (d, d') and 600'C (e, e').

crystalline bcc-Fe (H hf = 32.95 T, <5 = 0 mm/s) (Fig. 5. 9d- e). The spectral
contribution of this sextet increases with annealing temperature. The formation
of the bcc phase is accompanied by substantial changes in the retained
amorphous structure. The hyperfine field distribution consists now of two
distinct peaks. The main one is shifted to lower values of the hyperfine fields
(the average hyperfine field corresponding to this peak is 13.5 and 12.5 T for
the samples annealed at 550'C (Fig. 5.9d') and 600'C (Fig.5.ge'),
respectively, as compared to 16.2 T for the as-quenched alloy (Fig. 5. 9a') .
The intensity of the secondary peak in P (H), which appears at about 26.5 T,
increases with annealing temperature (Fig. 5. 9d' - e'). Such a shape of the
P (H) distribution suggests that Fe atoms encounter two distinct short range
orders in the remaining amorphous phase. The segregation of Fe atoms from
Mossbauer Spectroscopy Characterization of Soft Magnetic... 195

the original amorphous precursor and the progressive crystallization due to


increasing annealing temperature change the local composition in the
amorphous state, thus leading to the decrease of the average hyperfine field.
The 2 % Cu content in the FeZrBCu alloys dramatically promotes the
crystallization process. The crystallization temperature T Xl is decreased such
that annealing at 500'C causes substantial crystallization and the formation of
the bee-Fe phase (Fig. 5. 10c). The increase of the annealing temperature
results in a dramatic increase of the amount of the crystalline phases
(Fig. 5. 10d-e). The spectral component of bee-Fe dominates in the
spectra. The P (H) distribution reveals a complex shape with contribution
corresponding to the retained amorphous phase (with the peaks at hyperfine
fields smaller than 20 T whose relative contribution decreases with increasing
annealing temperature) and the peak at about 29 T corresponding most
probably to the interfacial regions (Fig. 5. 1Od' - e'). The great majority of Fe
atoms appear in the magnetic bee-Fe phase and the retained amorphous
phase. However, some Fe atoms experience zero magnetic fields, as
evidenced by the paramagnetic single line spectral component in the spectra
(Fig. 5. 10d - e), which is most probably related to the fcc-Fe phase. The
isomer shift of this line 5 = - 0.09 mm/s) agrees well with the isomer shift of
the y-Fe phase (Keune et al., 1977; Gonser and Wagner, 1985). The relative
amount of such a phase is too low to be clearly detected by the XRD
measurement.
Annealing at a temperature of 780'C causes complete crystallization of
the samples. Final crystallization of both alloys induces complete separation of
Fe atoms from other alloy components. The M6ssbauer spectra consist only of
one sextet characteristic for a-Fe.
Similar results were obtained for alloys with lower boron content (x = 8,
y=O, Fig. 5.11) and (x=8, y=2, Fig. 5.12) (Kopcewicz et aI., 1997a).
As can be seen from Figs. 5. lla and 5. 12a the spectra of the as-quenched
alloys are typical for the fully amorphous alloys and consist of the broadened
Zeeman sextet. The P (H) distributions, extracted from the spectra of the as-
quenched alloys, shown in Figs. 5. 11 a' and 5. 12a' , consist of a single broad
bell-like peak. After annealing at 430'C the samples are still completely
amorphous (Figs. 5. 11 band 5. 12b) .
As before, annealing at 500'C induces the onset of crystallization in both
alloys. The six-line spectral component with sharp lines, hyperfine field H hf of
32.9 T, and isomer shift 5 = O. 00 mm/s, all characteristic of the bee-Fe
phase, appears in the spectra in Figs. 5. 11 c and 5. 12c. The spectral
contribution of the bee-Fe sextet is much larger for the Fea3 Zr7 BaCU2 alloy than
for the Fea5 Zr7 Ba one. The P (H) distribution is only slightly affected by the
formation of the bee-Fe phase in the FeZrB alloy (Fig. 5. 11 c'). However, in
the Cu-containing alloy a second peak with <Hhf>~29 T appears in the P(H)
distribution, and the peak corresponding to the retained amorphous phase is
shifted towards higher H hl values (Fig. 5. 12c'). The change of the shape of
196 Michat Kopcewicz

1.00

0.95

1.00

"0
.;;:;
1.00
.'"
"'"
Cd
t:
(\) 0.95
>
.~
<;
0:: 1.00
4 .......... .e.
.. '... ....... (d')

0.90
1.00
o
4
........ ....
(e) .....' .. e.. ..... (e')

600'C
0.90

---{) 0 6 o 40
Velocity (mm/s)

Figure 5. 11 The Mossbauer spectra and corresponding P(H) distributions for the
FeS5Zr78S alloy in the as-quenched state (a) and (a'), and after annealing at 430 - 600 'c
(b) - (e) and (b') - (e').

the P(H) distribution reveals that the formation of the bee-Fe phase markedly
affects the retained amorphous structure. The iron content decreases and thus
the Curie temperature rises (Kobayashi et aI., 1986), leading to a larger
hyperfine field.
Annealing at temperatures of 550 'c (Figs. 5. 11d, 5.12d) and 600 'c
(Figs. 5. 11e, 5. 12e) drastically increases the spectral contributions
corresponding to the bee-Fe phase; the shape of the P ( H) distribution
changes markedly. The low field peak corresponding to the retained
amorphous phase is broadened, and the high field peak at <Hhf>~28 T related
to Fe atoms at the grain boundaries increases significantly (Figs. 5. 11 d', e'
and 5. 12d' , e'). With developing crystall ization, the abundance of the bee-Fe
grains and the grain boundary component increase, as seen in the M6ssbauer
Mossbauer Spectroscopy Characterization of Soft Magnetic... 197

.....
5 .
(a')

.. ...............
.....
0

5

(b')

0 .. .... ................
c
0
.;;;
.en
1.00
5
. .
en
g 0.95 (c) g .. (c')
..... .
Q)
>
.~

<>
"'-
.... ,"
......
0
~ 1.00 ......
. ...
4
.
... .. .. (d')

0.90
... '
.....
1.00
0
.... : .
4
.. . .
....... ...
(e) (e')

600C '
0.96
--6 0 6 0 20 40
Velocity (mm/s) Hhf(T)

Figure 5. 12 The M6ssbauer spectra and corresponding P (H) distributions for the
FeS3 Zr7 Bs OU2 alloy in the as-quenched state (a) and (a'), and after anneal ing at
430-6000 (b)-(e) and (b'-e').

spectra in Figs. 5. 11d, e and 5.12c-e.


Annealing the alloys at temperatures exceeding T X2 (Table 5. 1, in
Section 5. 3. 2. 1) causes complete crystallization of the amorphous phase.
The M6ssbauer spectra of all Fe93- x- y Zr7 B x CU y alloys annealed at T A =
780'C consist of a single spectral component characteristic of the ex-Fe phase. Thus,
the Fe atoms are almost completely separated from other alloy components.
The crystallization of alloys with x = 6 and y = 0 or 2 proceeds in almost
exactly the same way. The same bee-Fe phase appears at the same
temperatures, and its spectral contribution increases with increasing T A in a
similar way as in Figs. 5.11 and 5. 12.
The relative atomic fractions of the bee-Fe phase determined from the
relative area of the spectral component with the H hf = 32.95 T and {j =
198 Michat Kopcewicz

0.00 mm/s observed in the Mossbauer spectra of all alloy compositions vs.
annealing temperature are shown in Fig. 5. 13. These results are consistent
with the increase of boron content increasing the T Xl temperature, resulting in
the enhanced thermal stability of the amorphous phase. The relative
abundance of the bee-Fe phase increases with decreasing boron content for a
given T A above 430 'c (Fig. 5. 13a - d). The presence of 2 % Cu lowers the
T Xl temperature and dramatically promotes the precipitation of the bee-Fe
phase. This effect is seen especially for T A close to T x, , e. g., for T A =
500 'c (Fig. 5. 13b, c). The highest relative amount of the bee-Fe phase
formed at 600 'c is observed for low boron content (x = 6, Fig. 5. 13c) and at
650 'c for Fe89 Zr7 8 4 (Fig. 5. 13d) .

70 Fe79Zr7B,2Cu2
70 Feg3Zr7BgCU2
'+-
0 60 4-<
0 60
'"
<:
o Feg1Zr7B,2 <:
'V FegSZr7Bg
o~ o~
z?7- 50
<.>~
z?7- 50
u~

~'" '"~ 40
._ .J:
l:! '" 40
~ .J:
._ ~

~ i 30 ~ i 30
~ ~
';:' Q)
20 '">
''';::
'"'"(1) 20
~r.o.. ~U-

'"
0:: 10 '"
0::
10
As-quenched
0
o as-q
( I I I , , (
I I

400 500 600 400 500 600


Temperature ("C) Temperature ("C)
(a) (b)

70 70
FegSZr7B6Cu2 o Feg9Zr7B4
'+- '+-
0 60 o Feg7Zr7B6 0 60
<: <:
o~ o~
z?7-
<.> ~
50 ,z?7- 50
u~

~
~ '"
~ 40 ~
~ ~
._ .J:
'" 40
._ .J:

8"
"'.D
30 8"
"'.D
30
">
';:;
'"'" 20
Q)
"
> Q)
.~
'"'" 20
~r.o.. _u-
"
0:: 10 (\)
0::
10
As-quenched
0 0
, ( I I I ( I I I

400 500 600 400 500 600


Temperature ("C) Temperature ('C)
(c) (d)

Figure S. 13 The relative abundance of the bee-Fe phase determined from the
conventional M6ssbauer measurements.

All the alloys in the as-quenched state discussed above are ferromagnetic
at room temperature except for Fe89 Zr7 8 4 ; its Curie temperature is lower than
room temperature. For this alloy the ferromagnetic nanocrystalline bee-Fe
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 199

phase is formed in the paramagnetic amorphous matrix by annealing. Due to


the formation of the bee-Fe phase the content of Fe in the retained amorphous
phase decreases thereby causing the amorphous phase to become
magnetically ordered at room temperature. This effect is shown in Fig. 5. 14.
The Mossbauer spectrum of the as-quenched Fe89 Zr7 8 4 alloy consists almost
exclusively of the broadened quadrupole doublet (Fig. 5. 14a). Annealing at
T A = 475'0 causes the formation of the bee-Fe phase, as seen by the
appearance of the Zeeman sextet (Fig. 5. 14b) with a spectral contribution of
about 10%. This component drastically increases with increasing annealing
temperature (Fig. 5.14c-e).
1.001-_ _....

(a) As-quenched

0.80
1.00 ~oru-_~--v-:;_~-

(b)
0.90

0.95
650C

-{j 0 6
Velocity (mm/s)

Figure 5. 14 The M6ssbauer spectra and corresponding P(H) distributions for the
FeS9Zr7B, alloy in the as-quenched state (a) and after annealing at 475-650 C (b) - (e)
and (b')-(e').

There is an interesting change of the structure of the retained amorphous


phase upon annealing. At T A = 475'0 the central part of the spectrum
200 Michat Kopcewicz

broadens considerably and its shape changes. The amorphous matrix now
becomes magnetic and the hyperfine field distribution P ( H) extracted from the
spectrum in Fig. 5. 14b consists of a strong peak at low Hhf values H hf >""'"
3 T) and a smaller, much wider peak at about 13 T (Fig. 5. 14b'), suggesting
that the composition of the amorphous phase, is strongly inhomogeneous.
Higher annealing temperature strongly promotes crystallization and the spectral
contribution of the bee-Fe sextet increases to 53 %, 60 % and 65 % at T A =
500,550 and 650C, respectively (Fig. 5.14c-e and Fig. 5. 13d). The
P (H) distributions related to the amorphous matrix reveal the complex
structure. At T A = 500C (Fig. 5. 14c') the main peak in P (H) is shifted
towards larger Hhf and is observed at H hf ""'" 13 T. The second peak at H hf ""'"
31 T is most probably related to the grain boundary regions. The Fe content in
the amorphous phase is considerably decreased; the shape of the P (H)
distribution resembles those observed for the alloys with less boron discussed
above (Figs. 5.11 and 5. 12). The average hyperfine field in P(H) is related
to the iron content in the amorphous phase, and the lower Fe content with
increasing T A leads to higher Curie temperature (Kobayashi et aI., 1986) and,
hence, to larger values of H hf The contribution of the grain boundary
component increases with the increasing contribution of the bee-Fe phase, in
agreement with the increasing number of nanoparticles.
Taking into account the DSC and XRD results (see Sections 5.3.2. 1 and
5.8) we can expect that the bee-Fe phase formed at 500 C < T<600C is
nanocrystalline. However, the conventional Mossbauer study does not permit
verification of the grain size or the magnetic properties characteristic of the
nanoscale Fe grains, such as small anisotropy and coercivity and the vanishing
magnetostriction. Such measurements allow only the identification of the
phases in which iron is contained and the determination of the relative
abundance of iron-containing phases vs. annealing temperature and alloy
composition. It is also possible to follow the changes in the amorphous phase,
both before the onset of crystallization and after crystallization starts.
Other compositions of FeZrB (Cu) alloys have been also investigated. The
Fe86 CUI Zr7 B6 and Fe87 Zr7 B6 alloys have been studied by complementary
methods such as transmission electron microscopy, electron and X-ray
diffraction, resistometry, as well as the Mossbauer spectroscopy (Duhaj
et aI., 1996). Similarly to the earlier studies two steps in the crystallization of
the amorphous alloys induced by annealing were found. The first crystallization
begins with the formation of nanocrystalline ex-Fe. After the second
crystallization the nanocrystalline phase dissolves and together with the
remaining amorphous phase forms coarse grains of ex-Fe and dispersed Fe23 Zr6
phases. From the Mossbauer measurements two local environments of Fe
atoms were found in the amorphous phase: one, with low Zr content
responsible for the high-field component in P (H) distribution, and the second,
rich in Zr and B, corresponding to the low-field component in P ( H). Such
amorphous structure influences the subsequent crystallization process and is
Mossbauer Spectroscopy Characterization of Soft Magnetic... 201

the cause of different thermodynamical stability of the regions with different


local composition as indicated by the wide temperature interval between two
crystall ization stages.
An interesting phenomenon of magnetic hardening (increase of coercivity)
observed in the early stage of formation of the nanostructure was observed for
several alloys. XRD and M6ssbauer spectroscopy show that the magnetic
hardening is due to the appearance of a few nanocrystalls of Fe (e. g., in the
Fes72 Zr72 B43 CUll alloy) which are separated by a distance that is larger than
the exchange correlation length of the amorphous matrix. As the number of Fe
nanocrystals increases the intergranular distance decreases as a result of the
grains interacting by the exchange coupling giving rise to the subsequent
magnetic softening (Gomez-Polo et al., 1996). In order to explain this effect
the random anisotropy model (Alben et al., 1978; Herzer, 1991, 1993) was
modified to include a two phase nanocrystalline system (Hernando et al.,
1995) consisting of the nanocrystalline grains and the retained amorphous
matrix, in distinction to the original model which was developed for a single
phase magnetic system (Alben et al., 1978; Herzer, 1991, 1993). The
random anisotropy model explains the exchange interaction between the a-Fe
grains as the main origin of the soft magnetic properties of nanocrystalline
alloys. However, this model is only val id either when the sample is fully
crystall ized or when the exchange correlation length of amorphous matrix,
Lam' tends to infinity. For intermediate cases, e. g., in the early stage of
nanocrystallization or near the Curie point of the amorphous matrix (Lam"""'O),
the model should be modified by taking into account the two-phase nature of
the system.
Magnetic hardening is not only observed when nanograins are too far from
each other to interact by the exchange coupling but also when the retained
amorphous matrix becomes paramagnetic, i. e., at elevated temperatures.
Transition from ferromagnetic to paramagnetic state of amorphous matrix
causes a decoupling of the ferromagnetic grains and leads to the increase of
coercivity. The two-phase model was proposed by Hernando et al. (1995),
who considered the decay of the exchange interaction within the intergranular
amorphous regions. This model explains the magnetic hardening at the initial
stage of nanostructural evolution when the volume fraction of the crystalline
grains is low and when the volume fraction of the amorphous phase is
substantial. However, the magnetic hardening at temperatures near the Curie
point of the intergranular amorphous phase, at which the intergranular
exchange coupling is broken, could not be explained in terms of the Hernando
et al. (1995) model. This case was discussed by Suzuki and Cadogan
(1998), who proposed a new extended random anisotropy model, in which the
effect of the volume fraction and the exchange stiffness of the residual
amorphous phase are considered. The experimental M6ssbauer and coercivity
measurements performed for the nanocrystalline Fe9! Zr7 B2 alloy in the
temperature range from 77 to 4 73 K were interpreted successfully in terms of
202 Michat Kopcewicz

this new model, taking into account the effect of two local exchange stiffness
constants in the crystalline and amorphous regions on the exchange correlation
length (Suzuki and Cadogan, 1998).
The relationship between the mean hyperfine field of the residual
amorphous phase and the coercivity in various nanocrystalline samples has
been studied in order to explain the effect of the spontaneous magnetization in
the grain boundary amorphous regions on their magnetic softness (Suzuki and
Cadogan, 1999). Nanocrystall ine samples with various magnetization values
were prepared by anneal ing an amorphous Fe91 Zr7 B2 precursor for different
periods (60 s to 108 ks) at temperatures of 823 - 973 K. Coercivity clearly
decreased with increasing magnetization or decreasing volume fraction of the
residual amorphous matrix. This effect is well explained by the extended two-
phase anisotropy model (Suzuki and Cadogan, 1998) where both exchange
stiffness constant and volume fraction of the residual amorphous phase are
relevant to the exchange correlation length.
The magnetic hardening effect was observed also in other nanocrystalline
alloys, e. g., Feao,5Nb7B12,5 (Kopcewicz et al., 1999c). For various
nanocrystalline alloys the magnetic hardening near the Curie point of the
amorphous matrix was less prominent with decreasing volume fraction of the
amorphous phase related to the developing nanocrystallization of the alloys
(Hernando and Kulik 1994, Gomez-Polo et al., 1996).
The dependence of the Curie temperature and hyperfine field on tensile
stress was studied in amorphous FeZrB ( Cu) alloys (Barandiaran et al.,
1996). It was shown that amorphous FeZrB (Cu) alloys reveal a similar Invar
effect as the pure FeZr ones and that the Curie temperature strongly depends
on stress. The Invar effect depends on boron content and decreases with
increasing B-content. An applied tensile stress may induce paramagnetic to
ferromagnetic isothermal transition in some alloys. The effect of boron in Fe is
greater than that of Zr for enhancement of the ferromagnetism in these alloys;
it increases both the Curie temperatures and Fe magnetic moment, which is
related to larger electronic transfer to the 3d band of Fe.
In most of the papers discussed so far the Mossbauer spectra of
NANOPERM-type alloys were fitted using a superposition of one hyperfine field
distribution, P ( H), corresponding to the amorphous phase, and of a sextet
with narrow lines, with the hyperfine field of about 33 T and isomer shift {) =
0.00 mm/s, originating from the bee-Fe nanograins. Recently a modification of
this fitting model was proposed by Miglierini and Greneche (1997a, 1997b).
Their fitting procedure consists of considering two independent P (H)
distributions without assuming the shape of P (H) and, as before, one sextet
with narrow Lorentzian lines, corresponding to the bee-Fe nanograins. One
P (H) distribution with fairly low hyperfine fields is related to the retained
amorphous phase in the nanocrystalline alloy because the hyperfine field values
are close to those of the as-quenched starting amorphous precursors. The
slight asymmetry of the Mossbauer spectral component is taken into account by
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 203

assuming a linear correlation between the hyperfine field and isomer shift. This
is a standard assumption, commonly used earlier when fitting the M6ssbauer
spectra of conventional ferromagnetic amorphous alloys. At elevated
temperatures this P ( H) distribution can be decomposed to two components: a
magnetic component resulting from the ferromagnetically ordered amorphous
fraction and a non-magnetic component (quadrupole splitting) resulting from
the amorphous fraction in the paramagnetic state. The transition from the
ferromagnetic to the paramagnetic state of the residual amorphous phase is
rather broad in distinction to the as-quenched amorphous alloys. That feature
can be related to the structural and chemical inhomogeneities of the amorphous
matrix modified because of the formation of the bee-Fe grains. The second
P (H) distribution with larger hyperfine fields than in the first one is related to
the interfacial regions between the bee-Fe grains and the residual amorphous
matrix. This P ( H) block is most probably due to the outer iron surface of
crystalline grains. The low-field tail of the second P(H) distribution frequently
overlaps on the high field tail of the first P (H), what can be related to the Fe
atoms located in the amorphous phase but in contact with the bee-Fe grains.
The atoms located at the surface of the Fe grains and those located in the
amorphous phase in contact with the Fe crystals form the interfacial zone. An
example of the application of this method of fitting the M6ssbauer spectra
alloys is shown in Fig. 5. 15. This fitting method was successfully applied for
FesuZruB6Cu" FeaoTbB12Cul and FensNbuCrsCulB16 alloys (Miglierini and
Greneche, 1997a). A detailed analysis of the dependence of the P(H)
distributions on annealing conditions presented in the form of three-dimensional
mapping of the P (H) distributions, allowed the assignment of given parts of
the P (H) distribution to the amorphous and interfacial regions of
nanocrystalline alloys (Miglierini and Greneche, 1997b).
Similar processing of the M6ssbauer spectra of the nanocrystalline
Fen sNb u Crs CUI B 16 alloy including the 3D mapping of P ( H) distributions
exhibited the presence of atoms with different short range orders in the
amorphous and interface zone (Miglierini et a!., 1998). The paramagnetic,
weak magnetic and ferromagnetic regions were observed in the early stages of
nanocrystall ization.
The M6ssbauer results obtained for the same nanocrystalline FeNbCrCuB
alloy by Kopcewicz et a!. (1998b) were also processed using two P(H)
distributions which allowed an independent evaluation of the spin texture in the
retained amorphous phase and in the interfacial regions. It was found that upon
annealing the amorphous Fens Nb u Crs CUI B 16 alloy at a temperature of
450 'c, lower than the temperature of the first crystallization step, the
average hyperfine field at room temperature remains almost unchanged but the
spin orientation changes from random for the as-quenched alloy to the
preferential spin alignment in the plane of the annealed sample (Fig. 5.16) as
determined from the 0 23 parameter which changes from about 2 to about 4,
respectively. Annealing at T A ~ 490 'c, when the nanocrystalline phase is
204 Michat Kopcewicz

Crystal-amorphous Surface
interface of crystal grains

Interface zone

I Amorphous I Crystalline I

:y--
I

1.00

c
0 0.98
~
~

'~
c
g
Q)

.::::
:e 0.96
a:i
e:::

0.94 ~ --5 o
Velocity (111111/S)

Figure 5. 15 Schematic representation of a nanocrystalline structure and corresponding


Mossbauer spectra CMiglierini and Greneche, 1997a)

formed, as indicated by the recent DSC and XRD studies (Skorvanek et al.,
1997a, 1997b), caused clear changes in the Mbssbauer spectra (Fig. 5.16
c - e). The spectra of the two-phase system reveal the presence of three
components: G) the Zeeman sextet with a discrete value of the hyperfine field
Hhf~33 T and (j = 0.00 mm/s characteristic for the bee-Fe nanocrystals, and
two independent hyperfine field distributions, C2l one with the" small" field
Mossbauer Spectroscopy Characterization of 50ft Magnetic. . . 205

H hf =8 T and 0 23 =4 corresponding to the retained amorphous phase and the


other @ with the" high" field H hf =
29 T and 0 23 =
2 corresponding to the
interfacial region between the bee-Fe nanocrystals and the amorphous matrix.
In the crystalline/ amorphous interfacial regions Fe atoms located at the surface
of the nanograins and/or in their nearest neighborhood experience various
local atomic structures as a result of which the hyperfine field is reduced as
compared with Fe atoms in the core of the bee-Fe grains, which have a well
defined crystall ine structure.

1.00

5
0.95

1.00 o
5
0.95 (b)

o
1.00
"
0
.<;;
5
V>
6
V>

" 0.95
<l)
>
1.00
.~ 5
v~
(d)
0.95
o
VYT:rYV
~
1.00Wi"j"iW
5
0"95 (e)' f \V} :50~C
-6 0 6 o 10 20 30 40
Velocity (mm/s) Hhf(T)

Figure 5.16 The conventional M6ssbauer spectra recorded for the Fe73.sNb45CrsCu,B16
alloy in the as-quenched (a) and annealed (b) - (e) states. The corresponding magnetic
hyperfine field distributions extracted from the spectra (a) - (e) are shown in (a') - (e').
The spectral components related to the bee-Fe nanocrystals (i), retained amorphous
phase (ii) and interfacial regions (iii) used in the fitting procedure are shown in Fig. 5. le
as an example.
206 Michat Kopcewicz

The relative contributions of these spectral components change with


anneal ing temperature. The increase of T A causes an increase of the relative
amount of the nanocrystall ine phase from about 10% to about 20 % at T A =
490C (Fig. 5. 16c) and 550C (Fig. 5. 16e), respectively. The contribution
of the interfacial region also increases from about 17 % to 30 % at the expense
of the component corresponding to the amorphous matrix. It is worth noting
that spin orientations are different in the amorphous and interfacial regions.
There is a strong preference for the in-plane spin orienta,tion (0 23 ;:::,;:4) in the
retained amorphous phase, whereas spins of Fe atoms in the interfacial
regions reveal random orientation (0 23 ;:::,;: 2). The average hyperfine field at
room temperature corresponding to the amorphous phase decreased due to
annealing as compared with the as-quenched alloy (Fig. 5. 16a' -e'); that is,
related with a decrease of iron content in the retained amorphous phase.
These results are very similar to those obtained recently by Skorvanek et al.,
(1997a). This alloy was studied also using the rf-M6ssbauer technique
(Kopcewicz et al., 1998b) (see Section 5. 5. 2) .
Application of the two-distribution fitting model to iron based
nanocrystalline alloys was recently reviewed by Greneche and Miglierini
(1999) and Miglierini and Greneche (1999a).
Temperature dependence of the M6ssbauer spectra of some NANOPERM-
type alloys was investigated in order to understand better the intergrain
coupling in nanocrystalline alloys. The M6ssbauer measurements have been
performed for temperatures exceeding the Curie point of the residual
amorphous phase, i. e., the transition from the ferromagnetic to paramagnetic
state was observed. It was shown that the ferromagnetic coupling between
bee-Fe grains is mediated via about two-atom-layers thick interface regions
between nanocrystals (Kemeny et al., 1998). At temperatures well above the
Curie point, when the amorphous phase is in the paramagnetic state, the
nanograins can be treated as non interacting particles and may reveal
superparamagnetic relaxation (Kemeny et al., 1999). Superparamagnetism in
soft magnetic nanocrystall ine alloys was suggested also for
Fe66Cr6CulNb3Si,3B9 alloy (Slawska-Waniewska et ai., 1992). The
temperature dependence of amorphous-crystalline interface in Fes65 Zr65 B6CUI
(Miglierini et ai., 1997) and FesoNb7B12Cu, (Miglierini and Greneche, 1999b)
was investigated by two independent P ( H) distributions and 3D P ( H)
images. It was found that in early stages of crystallization small, a amount of
the bee-Fe nanocrystals are isolated from each other by the paramagnetic
amorphous residual phase. They affect the magnetic properties of the
amorphous phase creating chemical inhomogeneity that leads to the increase of
its Curie temperature. With proceeding crystallization the number of
nanocrystals increases, the intergranular distance becomes smaller and the
ferromagnetic exchange interaction between grains develops. A significant
role in this process is played by the amorphous-crystalline interfacial regions.
The transition to the paramagnetic state in the amorphous residual phase
Mossbauer Spectroscopy Characterization of 50ft Magnetic. . . 207

occurs in a wide temperature range; therefore it is not possible to characterize


this transition with a single Curie temperature. The role of the interfacial
regions was discussed also for the Fes5 Zr7 8 6 CU2 alloy (Slawska-Waniewska
et al., 1997; Slawska-Waniewska and Grenche, 1997). The thickness of the
interface was estimated from the M6ssbauer results to be about 1 - 2 atomic
layers.
Recently a class of NANOPERM-type alloys not containing Zr attracted a
lot of attention. A series of FeBoM78,2Cu, (M: Mo, Nb, Ti) alloys was studied
by complementary methods including DSC, XRD, TEM, STM and M6ssbauer
spectroscopy in the as-quenched amorphous state and after heat treatment
which induced the formation of the nanocrystalline phase (Miglierini et al.,
1999a). The magnetic hardening effect was observed in the early stage of
nanocrystallization in all these alloys (Fig. 5.17).
~

---+- Mo
90
____ Nb

_Ti
T 60
E
:5-
:J::u
30

o 300

Figure 5.17 Coercivity He at room temperature as a function of annealing temperature T A


for FeaD M7 8'2 Cu, (M = Mo, Nb and Tj) ribbons (Miglierini et ai., 1999a).

With increasing anneal ing temperature T A a magnetic softening occurs. In


the case of Mo-containing samples the lowest value of coercivity is found for
TA = 470 'C. The softening effect can be explained in the same way as by
Gomez-Polo et al. (1996), by the decrease of the intergranular distances due
to the increasing volume fraction of the nanocrystall ine phase as well as by the
changes in magnetic properties of the retained amorphous matrix, which allow
stronger exchange coupling between bcc-Fe nanocrystals.
The M6ssbauer measurements clearly show the influence of alloy
composition on the magnetic properties. The Mo-containing amorphous alloy is
paramagnetic at room temperature. The M6ssbauer spectrum of the
FesoNb7Cu,8'2 sample depicts broad and unresolved lines which indicate that
resonant atoms are already in a magnetically ordered state. The magnetic
ordering temperature of this sample is above room temperature (T c :::::::;333 K as
208 Michat Kopcewicz

obtained from magnetic measurements). Since the spectrum is quite broad, a


distribution of hyperfine magnetic fields P (H) was used to fit the experimental
data. The asymmetrical P (H) obtained, however, shows at least two peaks
located at about 5 and 12 T. The Feso Tb CUI B 12 amorphous alloy exhibits
stronger magnetic interactions at room temperature (T c ~ 413 K). The
corresponding M6ssbauer spectrum was evaluated by a P ( H) distribution
which shows an asymmetry towards low-field values indicating a presence of
magnetically different regions in the original amorphous phase.
M6ssbauer spectra of nanocrystallized samples consist of sharp Lorentzian
sextets which correspond to crystalline phases (CR), and of strongly
broadened sextets with distributed hyperfine fields which can be attributed to
the retained amorphous phase (AM). Iron atoms situated on the surface of
crystalline grains and those which originate from the amorphous matrix and are
in close contact with the crystalline grains form the interface zone (IF), which
is represented by the second P (H) distribution.
The use of two independent blocks of distributed parameters for the AM
and IF components allowed a more detailed study of the hyperfine interactions
and structural arrangements of the resonant atoms. Better qual itative and/or
quantitative assessment of changes in P (H) in the course of heat treatment
can be achieved with the help of 3D mappings of P(H) in Fig. 5.18.
The IF phase exhibits a stable short-range order arrangement featuring a
bimodal P(H)s for all compositions. The pronounced peaks in Fig. 5.18 can
be assigned to Fe atoms of the CR phase situated on the surface of the
crystalline grains whereas the smaller humps are attributed to Fe atoms placed
on the border between the CR and AM phases. No substantial variations in the
shape of 3D IF-P(H)s are observed. A decrease of the main IF-P(H) hump
for M = Ti is related to the fact that the CR phase grows more rapidly than for
M = Mo and Nb. Consequently, the nanocrystall ine grains are packed more
closely, thus effectively decreasing the number of Fe atoms located in the
outer surface of the CR phase. The disapearance of the low IF-P(H) hump for
440 and 470 'c annealed Ti-containing sample (Fig. 5. 18c right) is only
apparent because the respective H-values are included in the corresponding
AM-P(H) distributions (Fig. 5.18c left). In this particular sample it was
difficult to separate the contributions from AM and IF distribution blocks
because of rather high <H >AM values, i. e., pronounced magnetic interactions
within the AM phase.
As for the AM-P(H), the 3D images show a variety of hyperfine
interactions. Regions of different hyperfine interactions can be found in the
AM-P( H) suggesting a heterogeneity of the AM phase both from the
microstructure and magnetic points of view. Low hyperfine fields belong to the
AM atoms in non-magnetic positions. Rising importance of magnetic atoms is
demonstrated by a shift in the positions of P (H) humps towards higher
H-values. Their respective fractions, however, decrease as a consequence of
progressing crystall ization and a decrease in the AM content. Remarkable
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 209

Amorphous phase Interface zone

H(T) H(T)
(a)

P(H) P(H)

H(T) H(T)

(b)

P(H) P(H)

H(T) H(T)
(c)

Figure 5. 18 3D P (H) mappings for the AM and IF components of the


Feao M, 8'2 Cu, alloys for: (a) M = Mo; (b) Nb and (c) Ti (Miglierini et aI., 1999).

contribution of magnetic atoms is observed for M =


Mo (Fig. 5. l8c left) after
annealing. The as-quenched alloy was completely paramagnetic. One and two
210 Michat Kopcewicz

clearly distinct magnetic regions developed in AM of M =


Nb, and Ti,
respectively, as seen in the left-hand side of Fig. 5. 18b, c.
The formation of the nanostructure in Feao M78 12 CUI (M: Mo, Nb, Ti, Ta
and Zr) alloys and the temperature dependence of magnetic properties on the
composition of these alloys was studied by conventional and rf-Mossbauer
techniques (Kopcewicz et aI., 1999a; Grabias et al., 2000; Kopcewicz
et aI., 2002). Formation of the bcc-Fe phase proceeds in a similar way in all
these alloys. The relative atomic fraction of Fe as bcc phase vs. annealing
temperature is shown in Fig. 5.19 in which the results obtained earlier for the
Fe79 Zr7 8 12 CU2 alloy (Kopcewicz et aI., 1996) are included for comparison. As
can be seen, the increase of annealing temperature, T A' increases the
relative fraction of the bcc-Fe phase. At higher T A the relative fraction of bcc
Fe begins to saturate at around 50 %. The highest relative content of Fe
(56 %) was obtained for the Feao Th 8 12 CUj alloy annealed at 620 'c .
60
o M=Ti
M=Ta
'" M=Mo
M=Nb
o M=Zr

o 400 450 500 550 600 650


Anneal ing temperature (C )

Figure 5. 19 The relative abundance of the bee-Fe phase determined from the
conventional Mossbauer spectra of the Feso M7 8 12 CUI alloys.

The same alloys have also been studied by Miglierini and Greneche 1998.
The attention was focussed on the morphology of the interfacial regions in
FeM8Cu alloys using a 3D mapping of hyperfine field distributions, Which was
discussed earlier. Hyperfine fields of residual amorphous and interface phases
in the same group of nanocrystalline alloys were studied recently using the
Mossbauer technique at room and low temperatures (Miglierini and Greneche,
1999c; Miglierini et aI., 2001).
Structural and magnetic properties of the intergranular amorphous phase
in nanocrystalline Feao.5 Nb 7 8 12 .5 alloys was studied by Skorvanek et aI.,
( 2000 ) . The striking differences in the temperature dependence of
magnetization and the hyperfine field of the intergranular amorphous phase as
compared with the analogous amorphous alloy in the as-quenched state were
observed. These differences became particularly remarkable at temperatures
close to the Curie point of the amorphous phase. The results confirmed the
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 211

strong chemical inhomogeneity of the intergranular phase in FeNbB alloys.


Also in this alloy, the magnetic hardening effect at the early stage of the
nanocrystall ization process was observed. The conventional M6ssbauer
measurements revealed the formation of the bee-Fe phase with relative
spectral fraction increased substantially from about 11 % at T a = 490 'c to
50% at T a =610-650'C (Kopcewicz et a!., 1999c).
Soft magnetic properties of FeS6-xCulNbxB13(X=4, 5) alloys were
investigated recently by Girchardt et al. (1999).
Crystallization and magnetic properties of the Fes3BgNb7Cul alloy was
studied in a wide temperature range by Kim et al. (1996). It was observed
that the magnetization of the amorphous phase decreases more rapidly with
decreasing temperature than that of the nanocrystalline ferromagnetic phase
which may suggest the presence of the distribution of exchange interactions in
the amorphous phase or high metalloid content.
Formation of the nanostructure in the Feso Tb B 12 CUI alloy was studied by
transmission and conversion electron M6ssbauer spectroscopy, thus allowing
the comparison of the bulk and surface crystallization process (Grabias et a!.,
1999b). The relative abundance of the bee-Fe phase at the surface exceeds
those found in the bulk for all annealing temperatures.
Recently the first M6ssbauer measurements have been performed for the
HITPERM Fe44 C0 44 Zr7 B4CUI alloy (Kopcewicz et al., 2001) . The
nanocrystalline phase formed due to annealing at 650 'c was identified as the
FeCo phase.
The conventional M6ssbauer studies of various NANOPERM-type
nanocrystalline alloys may be summarized as follows:
(1) The crystallization of the amorphous precursor can be followed in
detail.
(2) The nanocrystalline phase can be identified as almost pure bee-Fe.
(3) The nanocrystalline phase can ~e distinguished from the retained
amorphous phase and interfacial regions.
(4) The relative fractions of the phases: nanograins, residual amorphous
phase and interfaces can be determined from the M6ssbauer spectra.
(5) The relative fraction of a given phase vs. annealing temperature can
be estimated.
(6) The evolution of magnetic properties of the amorphous phase can be
followed vs. annealing temperature and for various temperatures of the
measurement.
(7) The intergranular magnetic interaction through the amorphous phase
can be studied vs. temperature.
(8) The superparamagnetic behavior of nanograins can be observed.
However, the conventional M6ssbauer spectroscopy does not provide
information on:
( 1) magnetic anisotropy fields;
(2) coercivity;
212 Michat Kopcewicz

(3) magnetostriction;
(4) size of the grains.
Information regarding the magnetic anisotropy and magnetostriction in
nanocrystalline alloys may be obtained from the rf-Mossbauer technique which
will be discussed in Section 5.5.2, and the grain size should be determined by
other techniques (e. g., TEM, XRD, SAXS).

5.5.2 The rf-Mossbauer Studies

The unconventional technique which combines the Mossbauer effect with the
phenomena induced by an external radio-frequency magnetic field (rf collapse
and rf sideband effects) was applied for the first time for the study of
amorphous and nanocrystalline FeCuNbSiB alloys (Kopcewicz et al., 1994a).
The rf-Mossbauer experiment allowed us to distinguish the nanocrystalline
Fe3 Si phase from the microcrystalline phases formed in the course of
anneal ing. As discussed in Section 5. 2. 3, the rf-Mossbauer technique makes
use of the collapse of the magnetic hyperfine splitting due to fast magnetization
reversal induced by an external radio-frequency (rf) magnetic field. If the
frequency of the rf field is larger than the Larmor precession frequency and the
rf field is sufficiently strong to overcome local magnetic anisotropy, then the
magnetic hyperfine field is averaged to zero at the Mossbauer nuclei. The
rf-collapsed spectra, which appear in place of the magnetically split six-line
pattern, consist of a quadrupole doublet or a single line only, in spite of the
sample being in the ferromagnetic state. The rf-collapse effect is very
sensitive to even small changes of the magnetic anisotropy and occurs only in
soft ferromagnets. Thus, one can distinguish the very soft nanocrystall ine
phase from the magnetically harder microcrystalline one. The second rf-
induced effect, the rf sidebands, originates from the rf field induced vibrations
of atoms via magneto-acoustic coupling, which is magnetostriction. The rf
sidebands effect allows the study of magnetostriction properties, e. g., the
decrease of magnetostriction due to the formation of nanocrystalline grains.
The rf sidebands disappear when magnetostriction vanishes.
As an example of the application of the rf-Mossbauer technique, the
results obtained for the Fe8! Zr7 B 12 and Fe79 Zr7 B 12 CU2 alloys (Kopcewicz et ai.,
1996) will be discussed in detail. The Mossbauer spectra of these alloys in the
as-quenched and annealed state were measured during exposure to the
magnetic rf field of 20 Oe at 60.8 MHz. Typical results are shown in Figs.
5. 20a' - e' and 5. 21 a' - e'. For comparison the corresponding spectra
recorded without rf field are shown in Figs. 5. 20a - e and 5. 21 a - e. As can
be seen, a complete rf collapse of the magnetic hyperfine structure to a
quadrupole doublet is observed for both amorphous alloys (Figs. 5. 20a' and
5. 21 a'). The magnetic hyperfine field acting on the Mossbauer nuclei is
averaged to zero. The rf field applied is sufficiently strong to overcome the
Mossbauer Spectroscopy Characterization of Soft Magnetic... 213

magnetic anisotropy in the amorphous state and the fast magnetization reversal
is induced in response to the applied rf field. The hyperfine field vanishes
despite the fact that the sample remains in the ferromagnetic state, as
evidenced by the presence of rf sidebands in the rf collapsed spectra.
No rf field In rf field
1.00 1.00
\~~,..
'.J :: ;J
- (a')
0.95
0.96

1.00
\~
1.00

--- (b')

0.95
:j
-
r-
c
0
'v; 1.00
if>
.
: !
.--
if>
C
~ (c')
<l)
>
.~
0.90 :i
03
"
c.:: 1.00
,r-v-
- -
-.-,,.
- (d')
0.95 0.90

.' . . . . . . r-
1.00 1.00
:rf':!'flf\!'.
(e) : -. 1 , .. :: (e')
0.90
0.95 " ...
'. 1 i
--6 0 6 -6 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5. 20 The M6ssbauer spectra recorded for the as-quenched and annealed
Fe81 Zr7 B I2 alloy in the absence (a) - (e) and during rf exposure (a') - (e') to the field of
20 Oe at 60.8 MHz.

The characteristic features of the rf collapsed spectra remain almost


unchanged as long as the alloys are in the amorphous state, i. e., when the
annealing temperature does not exceed 500 or 470'C for FeZrB and FeZrBCu
alloys, respectively (Fig. 5. 20b'). The strong rf sidebands clearly seen in the
rf collapsed spectrum of the as-quenched FeZrB alloy (F ig. 5. 20a') markedly
decrease for the sample annealed at 500'C (Fig. 5. 20b'), suggesting that the
magnetostriction is reduced compared with that of the as-quenched alloy.
214 Michat Kopcewicz

No rf field In rf field
1.00
~~r~r
., ., (a')

'.
0.95 0.95 .'
.I
1.00
I

.r
.
-
. (b')
'.
0.96 0.90 :;
"
0 1.00 "0 1.00

..: ,~
Vi Vi
'"
.~
.!!!
E t
'"
~" "
"
.'::
~
"
.'::
.
.'
.:
(c')

;;; ;;;
-.; 0.96 -.; 0.95 'J
e<: e<: 1.00
1.00 ","'"
;fi/\...~~
~
(d')
0.98 ~
.:
;
0.95
,
':
1.00
: " !"\ ",,"!". r-
(e)
:=. -:.
. . ::.. ::.'
. ~:: .)
.. (e')
0.90 0.95

-6 0 6 -6 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5. 21 The Mbssbauer spectra recorded for the as-quenched and annealed
Fe79 Zr7 8 12 CU2 alloy in the absence (a) - (e) and during rf exposure (a') - (e') to the
field of 20 Oe at 60 8 MHz.

The formation of the bee-Fe phase causes a dramatic change of the shape
of the rf collapsed spectra. The central collapsed part consists now of a
superposition of a quadrupole doublet, corresponding to the retained
amorphous matrix, and a single Iine which corresponds to the magnetically soft
nanocrystalline bee-Fe phase. This effect is clearly seen in Figs. 5.20c' and
5.21 b'. The rf sidebands disappear completely in the case of the FeZrB alloy
annealed at 550 and 600'C (Fig. 5. 20c' - d') and are strongly reduced for the
FeZrBCu alloy annealed at 500'C (Fig. 5. 21 b'). This, together with the
appearance of a single line in the collapsed spectra, gives strong evidence
that the bee-Fe phase appears in the form of nanoscale, magnetically soft
grains and the entire alloy becomes zero magnetostrictive. However, for the
FeZrB alloy a noncollapsed component appears in addition to the fully
Mossbauer Spectroscopy Characterization of Soft Magnetic... 215

collapsed spectral component; this noncollapsed component consists of a


sextet with sharp lines, a hyperfine field and an isomer shift characteristic for
the ex-Fe phase (Fig. 5. 20c' - d'). This spectral component corresponds to
the ex-Fe phase formed due to annealing of the FeZrB alloy at 550 - 600C with
larger grains that have sufficiently high magnetic anisotropy to prevent the rf
collapse. Thus, by using the rf induced effects it is possible to distinguish the
magnetically soft nanoscale Fe grains from the magnetically harder
microcrystalline grains of ex-Fe. This is not possible in the conventional
Mbssbauer measurements in which all Fe grains give a common contribution to
the spectrum, regardless of their grain size and magnetic properties
(anisotropy, magnetostriction) (Fig. 5.20c-d). One can conclude that the
FeZrB alloy annealing at 550 - 600C causes the formation of Fe grains with
two quite different sizes: CD small ones, magnetically very soft, for which a
complete rf collapse was observed, and CV large ones with magnetic
anisotropy large enough to prevent any rf collapse. In the completely
crystallized FeZrB alloy (Fig. 5.20e) the rf effects cannot be induced. The
magnetic anisotropy of the microcrystalline Fe phase is sufficiently large to
prevent the external 20 Oe rf field from producing the reversal of
magnetization; the rf collapse is completely suppressed. The spectrum
recorded during rf exposure of the microcrystalline FeZrB sample
(Fig. 5. 20e') consists of a noncollapsed six-line component corresponding to
magnetically hard ex-Fe.
In the case of the Fe79 Zr7 B 12 CU2 alloy the size distribution of the bee-Fe
grains formed by annealing at 550 - 600 C differs considerably from that
observed for the FeZrB alloy. The increase of the annealing temperature
above 500C and the corresponding increase of the amount of bee-Fe grains,
observed in Fig. 5. 21c - d, causes a dramatic change in the rf collapsed
spectra (Fig. 5.21c'-d'). The component revealing the unresolved partially
collapsed hyperfine structure appears beside the fully collapsed components
(OS doublet and single line) (Fig. 5. 21c'). This is evidence of the formation
of grains with sufficiently large magnetic anisotropy to prevent the rf collapse
(noncollapsed component) in addition to the magnetically very soft
nanocrystalline grains of Fe for which a complete rf collapse is observed. The
shape of the noncollapsed component changes from partially collapsed (at
550C, Fig. 5.21 c') to one consisting of a broadened but already well
resolved sextet (at 600 C, Fig. 5.21 d'), suggesting that in this alloy
annealing causes the formation of grains with a broader distribution of
anisotropy fields as related to the size distribution of bee-Fe grains, embedded
in the amorphous matrix, in clear distinction to the FeZrB case discussed
above. More information regarding the distribution of anisotropy fields is
obtained from the study of the dependence of the rf collapse effect on the rf
field intensity, discussed below.
Annealing the FeZrBCu alloy at 780C causes a final crystallization, as
seen in Fig. 5. 21 e, and results in a complete suppression of the rf collapse
216 Michat Kopcewicz

(Fig. 5. 21e').
Let us discuss now the details regarding the shape of the central collapsed
part of the spectra recorded for the as-quenched and annealed at 550 - 600 'c
FeZrB alloy (Fig. 5. 20c' - d'). In order to increase the resolution of the
method. the rf collapsed spectra were measured with a reduced velocity
range. The details regarding the shape of the central rf-collapsed part of the
spectra recorded for the as-quenched and annealed at 600 'c samples of the
Fes 1Zr7 B 12 alloy can be found in Fig. 5. 22a. b, respectively. The spectrum in
Fig. 5. 22a, which consists of a well-resolved quadrupole doublet. was fitted
with two slightly broadened OS doublets (OS 1 = 0.62 mm/s and OS2 = 0.27
mm/s). These OS values correspond to the average values of two distinct
peaks in the P ( OS) distribution also fitted to the same spectrum. A similar
result has been obtained recently for the Fes6 Zr7 B6CUI alloy (Orue et al.,
1994). The shape of the rf collapsed spectrum changed considerably for the
sample annealed at 600 'c (Fig. 5. 22b) in which the nanocrystall ine bee-Fe
phase was formed. as evidenced in Fig. 5.20d. The spectrum in Fig. 5.22b
consists of three components: CDa fully rf-collapsed single line corresponding
to the nanocrystalline bee-Fe phase (cubic, therefore no quadrupole splitting;
isomer shift is the same as ex-Fe); (2) the quadrupole split component,
consisting of two OS doublets similar to those observed in Fig. 5. 22a.
corresponding to the retained amorphous matrix; @ two inner lines of the
noncollapsed sextet corresponding to the magnetically harder ex-Fe phase. The
single-line spectral component allows the determination of the relative
contribution of the magnetically soft nanocrystalline bee-Fe phase and to
distinguish it from the magnetically harder larger grains of ex-Fe. Since in the
spectra recorded without the rf field both nano- and microcrystalline Fe phases
reveal identical magnetic hyperfine splitting. it is not possible to distinguish
them in conventional Mossbauer measurements (e. g., Fig. 5. 20d, e). The
formation of the bee-phase affects the short range order in the parent
amorphous phase. As can be seen from Fig. 5.22 the relative spectral
contributions of both OS doublets are affected when the bee-phase is formed
(Fig. 5. 22a and b). The relative spectral contribution corresponding to the
doublet with OS = O. 27 mm/s decreases markedly in the annealed sample,
leaving another OS doublet with OS = O. 62 mm/s almost unchanged
(Fig. 5. 24b). This suggests that the nanocrystalline phase begins to form in
the amorphous matrix in the regions with the higher local symmetry (smaller
OS splitting) around the Fe atoms.
Similar results were obtained for other alloys from the Fe93-x- y Zr7 B x CUy
(x= 6, 8, 12; Y = 0, 2) alloys (Kopcewicz et al . 1997a). The results
obtained for the alloys with lower boron content (Fes5 Zr7 Bs and Fes3 Zr7 BsCU2)
are shown in Fig. 5. 23. The hyperfine split spectra of the alloys in the
amorphous state (Figs. 5. 11 a, band 5. 12a, b) collapse completely due to
the rf field of 20 Oe and the spectra consist of the quadrupole doublet
accompanied by rf sidebands (Fig. 5. 23a. b. a', b'). The sideband
Mossbauer Spectroscopy Characterization of Soft Magnetic... 217

1.0

(a)
As-quenched

<: 0.9
o
.iii
if>
.~
<:
,g 1.0
Q)
>
.~ (b)
~ 600'C

0.9

-3 -2 -1 0 1 2 3
Velocity (mm/s)

Figure S. 22 Comparison of the shape of the rf collapsed spectra of the amorphous


(5.24a) and nanocrystalline (5. 24b) Fesl Zr7 8 12 alloy.

intensities are smaller in the annealed samples as compared to the as-


quenched alloys (Fig. 5. 23a, b, a', b') suggesting that magnetostriction is
reduced due to anneal ing (Suzuki et a!., 1991 b). The rf sideband intensities
drastically decrease due to reduction of magnetostriction when the
nanocrystall ine phase is formed (F ig. 5. 23c, c').
The formation of the nanocrystalline bcc-Fe phase at T A = 500 'c
dramatically affects the shape of the collapsed spectra. The central collapsed
part consists of a superposition of a single line, corresponding to the cubic
nanocrystalline magnetically soft Fe phase, and a quadrupole doublet,
corresponding to the retained amorphous matrix (Fig. 5. 23c, c'). For
samples annealed at higher T A in addition to the completely collapsed central
part of the spectra, in which spectral contribution is strongly reduced, a well
resolved hyperfine structure appears (Fig. 5. 23d, e, e', d'). The clear
hyperfine-split pattern with markedly reduced average hyperfine field as
compared to ex-Fe observed in these alloys suggests that the magnetic
anisotropy in bcc-Fe grains is much smaller than in bulk ex-Fe, for which no
rf-induced narrowing is observed. Thus, the partially collapsed spectral
component corresponds to the nanoscale bcc-Fe grains having, however, a
magnetic anisotropy considerably larger than that for the amorphous phase,
whose hyperfine spl itting is completely collapsed by the rf field. The amount of
such grains markedly increases with increasing anneal ing temperature, as seen
by the dramatic increase of the magnetically split spectral component and the
decrease of the rf collapsed component at 550 'c (Fig. 5. 23d, d'). However,
218 Michal Kopcewicz

1.00 I~""","-"', r--v- 1.00


:r--v-'
, ,
(a) " As-quenched
, ' , ' (a')
0.90
" 0.90 "
f i
1.001-_ _.... .....
, r '
1.001-_-....
,
,
r
,
'
(b) , ' , ,'
430'C
(b')
0.90 0.90
:;
c
.~ 1.001---.... ,"'-......
_-
/
.~c (c)
': 500'C
g 0.90
<l)

.::: "

~
<;;
0.90 j
1.00
:1\ f: !'\ !'\
(dH ': : ' :' ~,
r 1.00

'~r-.ft: "f\r
" fA.. " :' '~ ' : ' : : ' ~, (d')
.- : 1 .... 0.98

0.95
) : f i
550'C
..;ew-..
" " '.J,.

. ~ :'
,~
\

1.00 ~ 1.00
tAl
(e) ~ ..
f\.('.
- !':...!1! "'\1', f., ,. f', f (e')
; ~ ,~
::~:~t~
ft: -. "
0.98 0.99

i 1'" '600'C
"=i~

-6 0 6 -6 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5.23 The Mbssbauer spectra recorded for the as-quenched and annealed Fee5 Zr, Be
(a) - (e) and Fe e3 Zr, Be CU2 (a' ) - (e') alloys during rf exposure to the rf field of 20 Oe at
60.8 MHz.

annealing at 600 'c causes partial restoration of the fully collapsed component.
thus revealing an increase of the relative abundance of the magnetically soft
nanoscale bee-Fe grains (Fig. 5. 23e, e'). Annealing at 600 'c seems to be
most favorable for the formation of the nanocrystalline Fe phase. Further
increases of annealing temperature (to 780 'C) cause complete crystallization
of the alloys and the formation of microcrystalline ex-Fe with magnetic
anisotropy large enough to prevent any rf-narrowing of the spectra.
The rf-M6ssbauer results show that annealing FeZrB (Cu) alloys at
temperatures 550 - 600 'c causes formation of grains with a fairly broad
distribution of the magnetic anisotropy fields related to a broad size distribution
of the bee-Fe grains embedded in the amorphous matrix. However, in alloys
with the highest boron content discussed above (Fesl Zr7 B 12 ) a bimodal grain
size distribution was detected by the rf-M6ssbauer technique. The presence of
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 219

Cu in the Fe79Zr7B12Cu2 alloy results in the formation of a more continuous grain


size distribution at 500 - 600 'c similar to that proposed to explain the data in
Fig. 5.23a-e and a'-e'.
Very similar results were obtained for the Fes7 Zr7 B6 and Fes5 Zr7 B6 CU2
alloys. The Fes7 Zr7 B6 alloy reveals the lowest Curie temperature and
magnetostriction in this set of samples, but in the rf collapsed spectrum of the
as-quenched alloy, the rf sidebands have been clearly detected. The spectra
of the samples annealed at higher temperatures are very similar to those
shown in Fig. 5. 23a - e for Fes5 Zr7 B s The Fes5 Zr7 B6 CU2 alloy behaves in the
same way as Fes3Zr7BsCu2 (Fig. 5. 23a' - e') but the relative contribution of
the fully collapsed component observed in the Fes5 Zr7 B6 CU2 sample annealed
at T A ?500 'c is somewhat larger than in the case of the Fes3Zr7BsCu2 one.
The splitting of the magnetic component related to the bee-Fe nanophase is
smaller than that observed in Fig. 5. 23d', e', but also remains quite well
resolved. Annealing at 780 'c causes complete crystallization of both alloys
with x = 6, which prevents rf collapse.
The rf collapse is very sensitive to even small changes of the ratio of the
effective anisotropy (which prevents the magnetization reversal) to the rf field
(which induces the magnetization reversal). Thus by changing the rf field
intensity, one can use the rf collapse effect to "scan" the distribution of the
anisotropy fields and to obtain more detailed information on the size
distribution of nanoparticles. The rf-Mossbauer measurements were performed
as a function of the intensity of the 60. 8 MHz rf field for all the as-quenched
alloys and those annealed at 500, 550 and 600 'C. Figure 5.24 shows, as an
example, the results obtained for the Fesl Zr7 B 12 alloy in the as-quenched state
(Fig. 5. 24a - e) and after anneal ing at 600 'c (Fig. 5. 24a' - e'). The rf field
intensity varied from 20 Oe down to 5 Oe. As can be seen, the amorphous
alloy is magnetically very soft; the complete rf collapse and rf sidebands are
observed for the rf field intensity exceeding 8 Oe (Fig. 5.24a-c), The rf
sideband, intensity decreases with decreasing rf field intensity, as expected
from the frequency modulation (FM) model of rf sidebands effect (Pfeiffer
et al., 1972). At 6 Oe the rf collapsed OS doublet starts to broaden (Fig.
5. 24d), and at 5 Oe the broadening is significant (Fig. 5. 24e). Such an rf
field intensity is comparable to the effective anisotropy field and is not
sufficient to induce a complete rf collapse. The broadening observed in
Fig. 5. 24e results from the partial restoration of the magnetic hyperfine
structure. For the FeS1Zr7B12 alloy annealed at 600 'c, in which the
nanocrystall ine bee-Fe phase was formed, the rf field intensity dependence of
the rf collapsed spectra reveals that the effective magnetic anisotropy
increased as compared to that of the as-quenched amorphous alloy. The
decrease of the rf field intensity from 20 Oe (Fig. 5. 24a', the same as Fig.
5. 20d', the details of which were discussed above in relation to Fig. 5. 22b)
to 16 Oe already broadens the central collapsed part. A further decrease of the
intensity of the rf field causes much stronger broadening of the entire central
220 Michat Kopcewicz

As-quenched
1.00 1.00 r - _....._
'.1~r-'r ~
(a)
'.
: 200e (a')
.

..
200e

0.90
:i
r-ou. __
t~
0.90 1.00
1.00

120e \! 160e
(b')
.
..
t: 0.90 t 6 0.90 :;
\\ Irr:
.~ 1.00
.~
t:
1_....,..._...
(c)
.r-v-
: : 80e
.~
'E
<I'>
t:
1.00

g
<l)
.: ,g (c') ,!
> ~ 0.95
.~ 0.90 .~

~ f v0::
1.00
\.:
1.00 I__~_" ,-......,-
~ ~"."
! 60e t f\ /I f
(d)
: :
.
. \ I! 10 Oe
.. (d')
,, .:

0.90

1.00 r-......."'"
(e)
F
\
'J
50e
0.95

1.00

(e')
~
; ~
.\
W

!
ri
~
6'oe

0.90 '-----'-----_ _L - -_ _L - - 0.95 ',,19/


'-----L--_ _L - -_ _L - -

-6 0 6 -6 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5. 24 The M6ssbauer spectra recorded as a function of the rf field intensity for the
as-quenched (a) - (e) and annealed at 600'C (a') - (e') FeS,Zr7B,2 alloy.

part of the spectrum and finally a well resolved magnetic hyperfine structure
appears (Fig. 5. 24c' - e'). All spectra of the annealed sample contain also
the magnetically split component, with the hyperfine field of about 33 T, which
corresponds to the relatively hard bee-Fe grains for which the narrowing due to
rf collapse does not occur, regardless of the rf field intensity (Fig. 5. 24a' -
e'). The magnetic anisotropy of these grains, which are most probably much
larger than those for which the rf field induces the collapse to a single line
(Fig. 5. 24a' , b), is much larger than the rf field intensity, thus preventing the
magnetization reversal. The rf field intensity dependence of the central
collapsed part of the spectra (Fig. 5. 24a' - e') reveals that the magnetic
anisotropy of small nanoscale Fe grains is larger than the effective anisotropy
in the starting amorphous alloy. Also the magnetic anisotropy of the retained
amorphous matrix in the annealed sample significantly increases due to the
Mossbauer Spectroscopy Characterization of Soft Magnetic... 221

formation of the nanocrystalline Fe phase, as compared to the as-quenched


amorphous alloy. This effect is clearly seen as a substantial broadening of the
rf collapsed component in Fig. 5. 24c' - e' , while in the similar rf field intensity
range the spectra of the as-quenched alloy revealed an almost complete rf
collapse (Fig. 5. 24c- d).
The increase of the magnetic anisotropy in the retained amorphous phase
in the annealed sample results from the change of the short range order which
is due to the change of local composition in the amorphous phase related to
partial phase separation and the crystallization of the nanoscale bee-Fe phase.
The formation of the nanocrystalline phase may induce local stresses due to a
volume difference between the nanocrystalline iron and the amorphous matrix
and, thus, increase the local anisotropy.
The same experiment performed for the Fe79 Zr7 B 12 CU2 alloy revealed
markedly different properties of the nanophase formed due to annealing. The
properties of the as-quenched alloy (Fig. 5. 25a - e) are almost identical to
those of the FeBl Zr7 B 12 alloy discussed above (Fig. 5. 24). The rf collapsed
spectrum of the sample annealed at 550C contains an unresolved, partially
collapsed component in addition to the central collapsed part consisting, as
before, of a quadrupole split part related to the retained amorphous phase,
and a single line corresponding to nanoscale bee-Fe (Fig. 5. 25a'). The
decrease of the rf field intensity to 16 Oe causes the appearance of a well
resolved hyperfine split partially collapsed component, leaving the central fully
collapsed part almost unaffected (Fig. 5. 25b'). With the further decrease of
the rf field intensity, the splitting of the magnetic hyperfine component
continuously increases and finally becomes close to that characteristic for the
bulk ex-Fe (Hhl~33 T) (Fig. 5. 25c' - e'). The spectral contribution of this
component increases substantially and dominates strongly in the spectra
recorded during rf exposure to the rf field of 10 Oe and less (Fig. 5. 25d' -
e'). The spectral contribution of the collapsed component, related to the
retained amorphous matrix, decreases simultaneously and contributes to the
broad unresolved spectral component. Since the rf collapse effect is very
sensitive to the local magnetic anisotropy, the increase of the splitting of the
hyperfine split component and the increase of its spectral contribution with
decreasing rf field intensity is related to the fact that the relation between the
rf field intensity and anisotropy field changes. For smaller rf fields the grains
with a given magnetic anisotropy, which at larger rf field intensity contributed
to the collapsed part of the spectrum, now contribute to the noncollapsed part.
Thus the spectra in Fig. 5. 25a' - e' provide evidence for a broad size
distribution of the bee-Fe grains. In this case, in clear distinction from the
FeBl Zr7 B 12 alloy (Fig. 5.24) the distribution of the grain size is fairly broad
and continuous. The magnetically hard, large Fe grains, observed for
FeBl Zr7 B 12 (noncollapsed component with H hI ~ 33 T), are not formed in
Fe79 Zr7 B 12 CU2 annealed at 550 - 600C (Fig. 5. 25a') .
222 Michal Kopcewicz

As-quenched 550'C
1.00
,f"\f'.;
., :. it
1.00
~~
(a) 200e (a')
.,, 200e
::

0.95
...
., 0.96
1.00
.c",,-
:.,
\I'\~r
1.00

(b) : 120e
(b')
V"" rv'\i .: 160e
0.95
.~
\1
c:
o
1.00
f 5 0.96

"II~ { ' r
Vi .~ 1.00
en
.~ Een .r'. A ,.. !'\
:..:' : \, ,.
r"
g (c)
:
80
e g
c:
(c') VI : : t Y
~ 0.95 '"
> ': 120e
.~ .~
-.; -; ~ 0.96 1
~
1.00
~
1.00
'1'\:.1\;
... .. ....f'\...~,.....r'"
.,,
1 60e
(d')
.; , 8 Oe
.
JI .. :.

, f ~~
0.95 V 0.96 f
1.00 r-..._...~ 1.00
,..""". 'I".
. :" 1".,,'" ,:
\. I 50e (e').: .: ~ : :. :.
\ ,I ,
.. .,,- ..
-y- ~ ..

0.95
V
L - L -_ _L -_ _L - 0.95
i ~ .'
L - L -_ _L -_ _L -
, 5 Oe

-6 0 6 -6 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5.25 The M6ssbauer spectra recorded as a function of the rf field intensity for the
as-quenched (a) - (e) and annealed at 550 'c (a') - (e') Fe79Zr7B12Cu2 alloy.

The measurements performed for Fe79 Zr7 B 12 CU2 annealed at 600 'c reveal
a very similar behavior, which occurs, however, at larger rf field intensities.
The shape of the spectrum observed in Fig. 5. 25d' for 8 Oe is similar to that
at 16 Oe for Fe79Zr7B12Cu2 annealed at 600 'C. Annealing Fe79Zr7B12Cu2 at 600
'c leads to a similar, fairly broad and continuous distribution of the Fe grain
size. However, the mean magnetic anisotropy of bcc grains formed in the
FeZrBCu alloy, is considerably smaller as revealed by rf-Mossbauer
measurements (Figs. 5.24 and 5.25) than that of the magnetically hard Fe
grains observed in the FeZrB alloy, which contribute to the magnetically spl it
component with Hhf~33 T in the rf collapsed spectra (Fig. 5. 20c', d'). Thus
the average size of the bcc grains in FeZrBCu alloy is markedly smaller than
the size of magnetically hard Fe grains formed in FeZrB alloy discussed above.
Another example is shown in Fig. 5.26 for the Fes3 Zr7 BsCU2 alloy. Again
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 223

the amorphous as-quenched alloy is magnetically very soft, so the decrease of


the rf field intensity from 20 Oe to 6 Oe does not markedly affect the shape of
the rf collapsed spectrum. The collapsed spectrum broadens significantly at
about 4 Oe (Fig. 5. 26e). This shows that in the case of the as-quenched alloy
the rf field of about 4 Oe is comparable to the effective anisotropy field which
starts to suppress the rf collapse effect. The rf sideband intensities decrease
with decreasing rf field intensity, as expected from the frequency modulation
model of rf sidebands effect (Pfeiffer et al., 1972; Kopcewicz, 1989). For the
sample annealed at 500C the rf-collapsed spectrum contains, in addition to
the fully collapsed central component, a partially resolved broadened six-l ine
component (Fig. 5. 26a'). The decrease of the rf field intensity from 20 Oe to
12 Oe does not significantly affect the central collapsed component. However,
the splitting in the hyperfine split component increases considerably
(Fig. 5. 26b', c'). At 6 Oe the central collapsed component broadens and its
spectral contribution decreases (Fig. 5. 26d'). Further decrease of the rf field
intensity results in a dramatic suppression of the rf collapse effect and in the
increase of the spectral contribution of the noncollapsed six-line component
whose splitting becomes similar to that characteristic of the microcrystalline ex-
Fe phase (Fig. 5. 26e'). A similar behavior is observed in Fig. 5. 26a" -e" for
the sample annealed at 550 C. Here, the spectrum recorded with the
strongest rf field appl ied (20 Oe, Fig. 5. 26a") consists of a central collapsed
component related to the retained amorphous phase and the dominating
magnetically split six-line component which is only slightly narrowed due to the
rf field. As before, a decrease of the rf field intensity causes a suppression of
the rf collapse effect and the central collapsed component disappears
(Fig. 5.26a"-d"). At 4 Oe the splitting of the noncollapsed component
approaches that characteristic of the ex-Fe phase (Fig. 5. 26e") .
Since the rf collapse effect is very sensitive to the local magnetic
anisotropy, the rf-Mossbauer spectra measured as a function of the rf field
intensity provide evidence for a distribution of anisotropy fields and, hence for
a size distribution of the bcc-Fe grains. They show that the bcc-Fe phase has
larger magnetic anisotropy than the amorphous precursor. The nanocrystalline
phase formed due to annealing at 500C is magnetically softer than that formed
at T A = 550C. As can be seen by comparing Figs. 5. 26a' - e' and 5. 26a"-
e" the collapsed pattern dominates in the spectra of the alloy annealed at 500
C (F ig. 5. 26a' - e') whi Ie in those recorded for the sample annealed at 550
C, the hyperfine split component is prevalent (Fig. 5. 26a" - e"). The
spectral contribution due to the rf collapsed component, seen in Fig. 5. 26a" ,
decreases markedly with the lowering of the rf field intensity and nearly
vanishes at H rf ""='6 Oe (Fig. 5. 26d"). Thus the bcc-Fe grains formed at T A =
550C reveal larger magnetic anisotropy and, hence have larger average size
than those formed at T A = 500 C. The distributions of the size of the bcc
grains formed at both T A temperatures are fairly broad and continuous as
suggested by the gradual changes in the shape of the rf spectra with rf field
224 Michat Kopcewicz

As-quenched 500"C 550'C


1.00

...r-v- ...."..,-- It,~ :\:" il'*


..

(a)
...: 200e :. " : :; .: (a")
ft
1
~
0.90
t
t

c 1.00
1.00

0.90
(b)
...r-v-
..
t c 0.90
(b')
.t.
..
~
~
160e

c 0.96
.1\1\
....
:.
,. , ..
I
....
1\ 1\ ",.
.. : ,......,(b")
.. tt;
.. , ....

. Mr
~

....f
.~ 1.00 .~ 1.00
0

ir::f
.;;;
'"
. '"
. '"
. :1\" .... ~
'"c '"c (c') '"c .. f#

'"
I:: 0.90
(c)
.. g 0.95 ..

g ,
: !

f '
,
(c")

"
> "> 5"
.~
I .~
"#
0.95

~ 1.00 ...... Q)
~'"
V":,"-"'",..
r'f\ .(tV-
lr 1.00 1.00
t
.....
0:

.- , ." ..
(d)
..
t l
, , . ,.
t
0.90

I
0.95
(d')
\iy 60e
0.95

, (d")

1.00 1.00

.....f .,"i ~ f1rf ~~!'r


... , ....
0.90
(e)
"I (e')
i
i'C 40e
0.95


, . ~aCe")
-6 0 6 -6 0 6 -6 0 6
Velocity (mm/s) Velocity (mm/s) Velocity (mm/s)

Figure 5.26 The Mossbauer spectra recorded as a function of the rf field intensity for the
as-quenched (a) - (e) and annealed at 500'C (a') - (e') and 550"C (a") - (e") Fea3Zrr
BaCU2 alloy.

intensity.
The Fes5 Zr7 B6 CUz alloy behaves in a similar way to all other Cu-containing
alloys in this set. The distribution of the anisotropy fields and, hence of the
bee-Fe grain size is continuous and more uniform than in the alloys without Cu.
The Mossbauer investigations of the Fe93- x- y Zr7 B x CUy alloys with and
without rf fields are summarized in Figs. 5.27 and 5.28 in which the typical
features obtained for the alloys annealed at 500 and 550 "C, respectively,
discussed in detail above, can be clearly seen. The most characteristic
features are the following:
(1) In all FeZrBCCu) alloys the bee-Fe phase is formed due to annealing
(Figs. 5. 27a - f and 5. 28a - f). This result is fully supported by the XRD
data. The increase of boron content in the alloys increases the crystallization
Mossbauer Spectroscopy Characterization of Soft Magnetic.. . 225

No rf field In rf field

~r-v-
1.00 1.00

...
.
(a')

:;
0.95 0.95

1.00
1.00 ..... "'.r-
. .
0.96
'.
"I
(b')

1.00
0.90
1.00 - '\r
.
.
45

"0
'Vi
"0
'Vi 0.90 (c')
OJ>
0.95 OJ>

'E 'E
OJ> OJ>
1
~" ~"
~~
1.00 1.00
"> ">
.~

v0:::
.~

v0:::
. =
0.95 '.
0.90 :j (d')
1.00
1.00
~r-
=
0.95 0.90 .
'. (e')

/
1.00 1.00

~rrv
..
0.90
(f) x=6
y=2
0.95
'."/
'.
"
(t')

-6 0 6 -6 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5.27 The M6ssbauer spectra recorded for the F6B3- x- y Zr, Bx CU y (x = 12,8,6; y =
0, 2) alloys annealed at 500'C in the absence (a) - (I) and during rf exposure (a')-
(f ') to the rf field of 20 Oe at 60. 8 MHz.

temperature resulting in the enhanced thermal stability of the amorphous phase


(Figs. 5. 27a, 5. 28a). The relative abundance of the bee-Fe phase increases
with decreasing boron content (Fig. 5.13). The presence of Cu in alloys with
the same boron content decreases the crystallization temperature and
dramatically promotes the precipitation of the bee-Fe phase (e. g.
Fig. 5. 27b, d, f);
(2) A complete rf collapse effect is observed only in fully amorphous
alloys (e. g., Fig. 5. 27 a'). The rf-Mossbauer spectra consist of the fully
collapsed component (qs doublet) accompanied by the rf sidebands; the rf
sidebands, clearly seen for the as-quenched amorphous alloys (F igs. 5. 20a' ,
5.21 a', 5. 23a and 5. 230, nearly disappear due to the formation of the
226 Michat Kopcewicz

No rf field In rf field
1.00 _OO=,,-,,-,r-,r-::"--
1,00~r----
,.
: : (a')
0.95 .'
0.90 '.j
1.00 1.00~~
....,
:
0.98
. ' (b')
0.95 '.j
1.00
1.00 ' " ",-
c c
.f\ " !'\ !'\ I
0 o .: :: :",; :: ;: (c')
.;;; .~
..
.. : ) ....
f i
'E'" 0.90 .~ 0.96 t
g'"c 1.00 g 1.00, ,;tW
o-"
>
ve<:
~
.~
~
,'~I#,
....."'!\'
.: .;
.,
: :~.
.::
O98
. :.:) 'J.
. ~i
i.,' (d')
0.90
1.00
1.00 " " ~
!I\f\ ,,-A' (e')
::.,: .:!\:
s-
O 98 , ' .'
0.90 . 'i-'I\(
, '1
1.00

if\.,~
1.00
!'\!'f
'oj :. .: 'i
x=6 ( :;' (f')
0.95 (f)
y=2 'oj
0.96 L-':-_----'-!,-----_---':-_
-6 0 6 -6 o 6
Velocity (mm/s) Velocity (mm/s)

Figure 5. 28 The M6ssbauer spectra recorded for the Fe.13 - x - yZr7 Bx CU y (x = 12, 8, 6; y =
0, 2) alloys annealed at 550'C in the absence (a) - (f) and during rf exposure (a') -
(f ') to the rf field of 20 Oe at 60. 8 MHz.

nanocrystalline phase. This is related to the reduction of magnetostriction;


(3) The onset of the bee-Fe phase affects the shape of the rf collapsed
spectra; in addition to the OS doublet, a single line appears. The
noncollapsed six-line spectral component (H hf about 33 T) appears in the rf
spectra of alloys without Cu (Figs. 5. 27c' , e' and 5. 28a') implying that a
bimodal distribution of the bee-Fe grain size appears at the early stage of
crystall ization;
(4) Addition of Cu leads to a more homogeneous size distribution of bcc-
Fe grains. However, the anisotropy of the nanocrystals is higher than that of
the amorphous phase. The rf collapse is not complete so the partly narrowed
magnetic hyperfine split pattern appears in the spectra recorded during rf
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 227

exposure (Figs. 5. 27d' , f ' and 5. 28b' - f ') ;


(5) The spectral contribution of the partly narrowed hyperfine structure of
the nanocrystalline bee-Fe phase increases, at a given annealing temperature,
with decreasing boron content in Cu-containing alloys (Figs. 5. 27d', f 'and
5. 28b', d', f ') ;
(6) The nanoscale bee-Fe grains are smaller (lower anisotropy) in the
Cu-containing alloys with the same boron content, e. g., Figs. 5. 28c', d and
5. 28e' , f', which show larger spectral contribution of the rf collapsed component
(despite the fact that the abundance of the retained amorphous phase is lower)
and stronger narrowing of the noncollapsed hyperfine split component.
The rf-M6ssbauer technique was used recently for studying the magnetic
properties of the Fens Nb 4 Crs CUI B 16 alloy in the as-quenched state and after
thermal treatment, a result of which is the soft magnetic nanostructure was
formed (Kopcewicz et al., 1998b). Similarly to the FeZrBCu alloys, also in
the latter alloy the rf-M6ssbauer measurements permitted evaluation of the
magnetic anisotropy fields. The magnetic anisotropy of the nanocrystalline
bee-Fe phase was found to be considerably larger than that corresponding to
the amorphous phase. The rf sidebands appear only in the fully amorphous as-
quenched alloy. The sidebands disappear in the annealed alloys which is
related to the reduction of magnetostriction due to the formation of the
nanocrystall ine phase.
Similar effects have been observed for the Fesos Nb 7 B I2 .S nanocrystalline
alloy (Kopcewicz et aI., 1999c). Also in this case the complete rf collapse of
the magnetic hyperfine structure was observed for the original amorphous
phase and for the residual amorphous matrix in the nanocrystalline alloy. The rf
sidebands, observed in the rf-M6ssbauer spectrum of the as-quenched
amorphous precursor, disappeared when the nanocrystalline bee-Fe phase was
formed upon annealing. The rf-M6ssbauer spectra recorded for the
nanocrystalline alloy consist of two components: (Da rf-collapsed quadrupole
split doublet originating from the retained amorphous phase, and cg)a sextet
corresponding to the bee-Fe grains. The collapsed doublet is markedly
broadened for the annealed samples as compared to that recorded for the
as-quenched alloy, which suggests that the anisotropy field in the retained
amorphous phase is larger than that in the as-quenched alloy. The reason for
this could be the stress induced by the formation of the bee grains in the
amorphous matrix or the change of its composition towards the iron-poorer
one. However, at higher annealing temperatures the rf-collapsed doublet
narrows again, revealing that the anisotropy in the retained amorphous phase
decreases at higher T A. The bee-Fe sextet is only marginally narrowed by the
rf field indicating that the anisotropy in the bee-Fe grains is considerably larger
than that in the retained amorphous phase. The magnetic anisotropy in the
nanograins is, however, considerably smaller than in the microcrystall ine
ex-Fe. These conclusions were confirmed by a detailed rf-M6ssbauer
measurements as a function of the rf field intensity.
228 Michat Kopcewicz

In addition to the Mossbauer study, the magnetic measurements,


performed for the nanocrystalline FeNbB alloy in various stages of
nanocrystallization, revealed that the coercive field, He' depends strongly on
anneal ing temperature. The smallest H e ~ O. 025 Oe was observed for the
amorphous alloy. Annealing at 490 - 530 'c increased the coercivity by two
orders of magnitude due to the reduced strength of the exchange interaction
between the adjacent bee-Fe grains separated by a relatively thick weakly
magnetic amorphous matrix in the alloy with a fairly small fraction of
nanocrystals. Coercivity reaches a maximum of 2.6 Oe for T A = 530 'C. At
higher annealing temperatures (570 - 610 'c ) the volume fraction of the bee-Fe
grains strongly increases, which leads to a more effective intergranular
interaction and results in the strong reduction of the coercivity. The minimum
He = O. 15 Oe was observed for T A = 610 'c . Our rf-Mossbauer measurements,
which are sensitive to the anisotropy fields rather than to the coercivity, reveal
the strongest narrowing of the magnetic hyperfine structure (mhfs) for the
sample with the smallest He' thus suggesting that the best soft magnetic
properties are obtained in this alloy annealed at 610 'c .
Influence of the alloy composition on the magnetic properties of
nanocrystalline FeBOM7B12CUl (M: Ti, Ta, Nb, Mo) alloys was studied recently
by the rf-Mossbauer technique (Kopcewicz et ai., 1999a and 2002; Grabias et
al., 1999b, 2000; Miglierini et ai., 1999). In the Ti-, Ta- and Nb-containing
alloys the rf-induced effects occur in a similar way, revealing similar behavior
of the rf-Mossbauer spectra. As an example, the case of Feso Tiy B 12 CUI alloy is
discussed. Typical rf-Mossbauer spectra measured during exposure to the rf
field of 12 Oe at 61. 2 MHz are shown in Fig. 5. 29. The hyperfine spl it
spectrum of the alloy in the amorphous state (Fig. 5. 29a) collapses
completely to the quadrupole split doublet (OS) accompanied by the rf
sidebands, thus revealing that the applied 12 Oe rf field is sufficiently strong to
overcome the magnetic anisotropy in the amorphous state. The spectrum
recorded for the two-phase nanocrystalline alloy (bee-Fe embedded in the
retained amorphous matrix) consists, similarly to the case of the FeZrBCu
alloys, of two components (Fig. 5. 29b' - e'): CD a fully collapsed central
( OS) doublet originating from the retained amorphous phase and ~a partially
collapsed six-line component related to the nanocrystalline bee-Fe phase. With
increasing annealing temperature the partially collapsed component becomes
better resolved. Even though the hyperfine splitting of this component for the
sample annealed at 620 'c is the largest in the set of the spectra in
Fig. 5. 29b' - e', the average hyperfine field is markedly reduced as
compared to bulk ex-Fe as well as to the bee-Fe component in the spectrum
recorded for this sample in the absence of the rf field (Fig. 5. 2ge). The
reduction of the hyperfine field due to the rf field strongly suggests that the
magnetic anisotropy in the bee-Fe nanograins is considerably smaller than in
the bulk ex-Fe but is sufficiently large to limit the fast magnetization reversal
and is considerably larger than that in the retained amorphous phase for which
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 229

a complete rf-collapse is observed. The spectral fraction of the partially


collapsed six-line component increases with increasing annealing temperature
(Fig. 5. 29b' - e') which is related to the increase of the amount of the bcc-Fe
nanograins at the expense of the amorphous phase (Fig. 5.19). The observed
changes in Fig. 5. 29a' - e' in the rf-M6ssbauer spectra are characteristic of
all nanocrystalline alloys studied.
No rf field In rf field

....,.-...,-
1.00

(a')
0.90

0.80
,..
1.00

:i
N'J (b')
\:
0.95
.'i
c: c:
0
'Vi .~ 1.00
,.."-
'"
.~
c:
~
.~

~
c: ......
~
'
f4.,,,\ 1':. \:
., . . \,... \
'
(c')

.,
'"
.~
5'"
.; l

"
.... ",
~
0.96
Q) '"
Q)

'" '" 1.00

0.98
:"" "At
': ::' .,' i. (d')
....
.'
~
. ..'.-
,
..,.,
f
':II
1.00
r
If~
: ." ' ....:
,I\!\
( ')
:: .: : : :::: e
...Lt ....
(e)
0.96 ' ' ' , ... 1 "
0.90 '------'-----~_ _'__~_-'-_
--{i 0 6 --{i 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5. 29 Mbssbauer spectra of the as-quenched and annealed Feao Tb 8'2 Cu, alloys
recorded without Ca) - CO, and with rf field of 12 Oe Ca') - (f ').

The rf-M6ssbauer measurements performed as a function of the rf-field


intensity closely resemble those obtained for the FeZrB(Cu) alloy (Fig. 5.28)
and provide information regarding anisotropy fields in each phase present in
the nanocrystalline alloy. Also in this case the influence of the rf field intensity
on the rf collapse in the as-quenched and annealed alloy was studied.
The anisotropy fields in nanocrystall ine phases formed at different
annealing temperatures can be compared through the degree of narrowing of
230 Michal Kopcewicz

the partially collapsed six-I ine spectral component related to nanograins. The
least rf-narrowed magnetic component was observed for the FeTiBCu alloy
annealed at 440 'c (the average hyperfine field is about 28 T, Fig. 5. 29b' ) .
This value decreases to about 25 T for T A = 470 'c (F ig. 5. 29c') and
stabilizes at about 26 T for TA~520 'c (Fig. 5.29d', e'). This suggests that
the anisotropy in the bee-Fe grains formed at the early stage of crystallization
is somewhat larger than that characteristic for the grains formed at TA~470
'C. These observations agree well with the measurements of the coercivity
vs. T A for this alloy (Miglierini et aI., 1999) which increases strongly at low
T A (440 - 470 'C) and dramatically decreases for higher annealing
temperatures (Fig. 5.17).
The rf-M6ssbauer spectra recorded for alloys with different M substitutions
were compared. As an example the spectra of the samples annealed at similar
temperatures and recorded during the exposure to the rf field of the same
intensity (16 Oe) are shown in Fig. 5. 30. The spectra recorded in the
absence of the rf field (Fig. 5. 30a - e) reveal similar relative fractions of the
bee-Fe phase in all alloys. However, the rf-M6ssbauer spectra differ
considerably. The most effective collapse of the magnetic hyperfine structure
was observed in Fig. 5. 30b', c' for Nb- and Mo-containing alloys,
respectively, and in Fig. 5. 30a' for the FeZrBCu alloy. The fully collapsed
central part dominates in these spectra and the partially collapsed six-line
component is narrowed most. This suggests that the alloys containing Nb, Mo
or Zr are magnetically the softest, both in regard to the retained amorphous
phase and the nanocrystalline bee-Fe grains. The spectra of the Ti-and Ta-
containing alloys (Fig. 5. 30d', e') are much less narrow. The spectral
contribution of the central collapsed part is much smaller than in the case of
Nb- and Mo-containing alloys. This suggests that the magnetic anisotropy is
considerably larger in Ti- and Ta-containing alloys due to larger (as revealed
by XRD measurements, see Section 5. 8) bee-Fe grains as compared with
Nb-, Mo- and Zr-containing alloys. Similar characteristic features were
observed for the samples annealed at lower temperatures.
Recently the first rf-M6ssbauer results were obtained for the HITPERM
(Fe44 C0 44 Zr7 B4CUI) alloy (Kopcewicz et aI., 2001). The rf-M6ssbauer results
have shown that the nanocrystalline FeCo phase has a fairly large magnetic
anisotropy, certainly larger than in NANOPERM or FINEMET alloys. However,
a small fraction of the nanocrystalline grains are very soft and rf-collapse was
possible to induce. The larger anisotropy of the FeCo phase is expected for
these alloys due to the compositional (magnetostriction and magnetocrystalline
anisotropy) and microstructural (slightly larger size of nanocrystallites)
effects. By optimization of these parameters, the alloy can be tuned for better
performance.
Our discussion of the application of the rf-M6ssbauer technique to the
characterization of the nanocrystalline alloys can be concluded as follows. The
unconventional rf-M6ssbauer technique, in which the rf collapse and sideband
Mossbauer Spectroscopy Characterization of Soft Magnetic... 231

No rf field In rf field (16 Oe)

~~
1.00 1.00

V\. M=Zr
0.98
..
if 500C
I
1.00
0.96 V I

.~
1.00
~

0.95
0.98 : .
: M=Nb
580C
,
I
c: 'J
~
c:
0
'<ij
1.00 0 1.00
.~

i \ v: .
'" .
.~
c: '"c: 1\" !'IV I
I
:.
(c')
g g M=Mo

.~>'"
.,
'">
.,
.~ 0.96 I
I
I
I .
\j
570C
I
I
~ ~ 1.00 I ~
1\ ,.
(d') I. :'\
I ,
I
. I... ...
. i ."'f\:
I I
Ir
M=TI

I :: t 580C
0.98 II
It"
-It: "I
I
I
1.00

I "t)
4
~r~/, Ifrt
(e) (e')I.:t 'i : : ; I
I. . I ~. I
I' : 'M=Ta
0.95 0.98 Ii
I:
r-.. '580 C

--6 0 6 --6 0 6
Velocity (mm/s) Velocity (mm/s)

Figure 5. 30 M6ssbauer spectra of the annealed Feso M) 8 12 CUI alloys recorded without
Ca) - Ce), and with rf field of 16 Oe Ca') - Ce').

effects are observed, permits us to distinguish the magnetically soft


nanocrystalline bee-Fe phase from the magnetically harder microcrystalline
a-Fe and provides information concerning the distribution of anisotropy fields
related to the distribution of the size of the bee-Fe grains. Thus, this technique
is superior to the conventional M6ssbauer experiments which allow only the
identification and determination of the relative abundance of the phases
present in the material.
Qualitative information concerning the distribution of anisotropy fields
related to the distribution of the size of the bee-Fe grains can be inferred from
the dependence of the rf collapsed spectra on the rf field intensity. The rf-
M6ssbauer results show that the nanocrystalline bee-Fe phase, being
magnetically very soft, has markedly larger anisotropy than that of the parent
amorphous phase. The rf-M6ssbauer measurements provide information
232 Michat Kopcewicz

regarding magnetic anisotropy in each phase present in the sample in clear


distinction to the conventional magnetic measurements in which the coercive
field is measured for a whole composite sample. The rf-Mbssbauer technique
delivers a unique set of information on the microstructure and magnetic
properties, particularly when combined with other techniques, such as XRD
and SAXS, which allow direct estimation of the size of the nanoscale bcc-Fe
grains.

5.6 Surface and Bulk Nanocrystallization of Amorphous


FeCuNbSiB and Fe-M-B-Cu Alloys-the CEMS Study

It is well known that the ribbons of amorphous alloys prepared by the melt-
spinning technique have two distinct surfaces: a dull one (which was in contact
with the quenching wheel) and a shiny one (free surface), whose atomic
compositions may differ somewhat from that of the bulk. Therefore the onset of
crystallization induced by annealing may be significantly affected by the local
composition fluctuations and may be different at the surfaces and in the bulk of
the ribbon. As discussed in Sections 5. 4 and 5. 5, the Mbssbauer
spectroscopy is a powerful technique for studying the crystallization process of
amorphous alloys. However, in the Mbssbauer studies discussed above, only
the transmission geometry was used, which allowed the study of the
crystallization process in the entire volume of the sample (bulk
crystallization). Earlier Mbssbauer studies of the crystallization process in
amorphous FeB, FeP (Wagner et aI., 1985), FePC (Gonser et aI., 1983),
FeBSi (Saegusa et aI., 1982; Ok and Morrish, 1981), FeCoB (Zemcik et aI.,
1993) revealed distinct differences between bulk crystall ization and that
occurring at the sample surfaces. Conversion electron Mbssbauer
spectroscopy (CEMS), probing the surface regions about 120 nm thick, is an
ideal method for studying the surface crystallization. Combination of CEMS and
y-transmission Mbssbauer measurements allows simultaneous observation of
the effects induced by annealing at the surface and in the bulk of the sample. In
this way the surface and bulk crystallization behaviors can be readily
compared.
Preliminary CEMS results obtained for Fe735 CUI Nb 3 Si,65 B6 (Jiang et aI.,
1991) and Fe735 Cu, Nb 3 Si 135 B9 (Pundt et al., 1992) already some differences
between the bulk and surface spectra. In the first case small differences in the
spin texture at the surface as compared with the bulk were detected for the
sample annealed at 820 K. Pundt et al. (1992) noticed that for FINEMET
sample annealed at 520C for 2 min the surface shows sharp lines in the CEMS
spectrum indicating crystallization while the bulk of the sample still shows only
the amorphous structure.
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 233

Systematic CEMS studies of the NANOPERM-type alloys revealed clear


differences between the surface and bulk crystallization of an amorphous
precursor in which the nanocrystalline phase was formed by annealing
(Kopcewicz et al., 1996, 1997a). As an example the results obtained for
Fesl Zr7 B I2 and Fe79 Zr7 B I2 CU2 are discussed in more detail (Kopcewicz and
Grabias, 1996; Kopcewicz et ai., 1997c; Grabias and Kopcewicz, 1998).
As can be seen from Fig. 5.31, the transmission and CEMS spectra of the
as-quenched Fesl Zr7 B I2 alloy are similar. They consist of a broadened sextet
due to hyperfine field distribution characteristic for a purely amorphous state.
However, clear differences between the bulk and surface crystallization
appear already at T A = 430 C. The transmission spectrum (Fig. 5. 31 b)
consists of the spectral component characteristic for the purely amorphous
state in which no crystalline phases could be detected, while the CEMS
spectrum (Fig. 5. 31 b') reveals, beside that component, the magnetically
split spectral component with the hyperfine field of about 33 T and 0 =
0.00 mm/s characteristic for the bcc-Fe phase. Thus, the crystallization
process started already at T A = 430C at the shiny surface of the sample. In
the CEMS spectrum recorded for the dull side, only traces of the bcc-Fe phase
could be detected. A similar situation was observed for the sample annealed
at 500C (Fig. 5. 31 c and c'). The spectral contribution in the CEMS
spectrum of the sextet characteristic for bcc-Fe increases markedly with
increasing annealing temperature. At T A = 550 C, which exceeds the T x,
temperature, the bcc-Fe phase appears in the bulk (Fig. 5.31 d). The relative
fraction of this phase' is, however, much lower than that at the surface (Fig.
5.31 d') .
In the CEMS spectra of the Fesl Zr7 B I2 alloy annealed at T A = 430, 500
and 550 'c beside the bcc-Fe sextet, a paramagnetic component, not
observed in any transmission spectra, consisting of the quadrupole splitting
( OS) doublet (OS ~ O. 80 mm/sand 0 ~ O. 34 mm/s ) is clearly seen
(Fig. 5.31b'-d'). The origin of this component is uncertain. The spectral
parameters of this doublet do not correspond to any of the crystall ine FeZrB,
FeB or FeZrO phases. However, the observed OS and 0 values correspond
quite well to that characteristic for the iron-poor paramagnetic amorphous FeB
phase (Hoving et al., 1985). Such a phase could appear at the surface as a
result of segregation of some iron atoms from the amorphous matrix to form
the bcc-Fe phase. The amorphous FeB phase is unstable and decomposes
with progressing crystallization, which occurs at higher annealing
temperatures.
The CEMS spectrum of the Fesl Zr7 B I2 sample annealed at T A = 600 C
(Fig. 5. 31 e') contains, in addition to the sextet characteristic of the bcc-Fe
phase, very intense spectral components (H hfl = 23.2 T and H hl2 = 24.2 T and
01 = O. 12 mm/sand 02 = O. 11 mm/s, respectively) corresponding to the
tetragonal Fe2B phase <Takacs et ai., 1975) which contribute to about 40% of
the total spectral area. This phase is formed only at the surface of the sample
(on both shiny and dull sides). The transmission spectrum in Fig. 5. 31e does
234 Michat Kopcewicz

Transmission CEMS
1.00 1.01

0.96 1.00
1.01
1.00

1.00
1.01
c c
0 0
'Vi 'Vi
'6'" '6'" 1.00
'"c 0.96 ,.'"c
g 1.00 t: 1.01
<l) <l)
> (d) >
~
.~

v
~ 0::
0.96 1.00
1.00
1.01

1.00
1.02

0.90
780'C
1.00
--6 -3 0 3 6 --6-3036
Velocity (mmls) Velocity (mm/s)

Figure 5.31 Mbssbauer transmission (a) - (f) and corresponding OEMS (a') - (n
spectra recorded for the as-quenched Fesl Zr7 8 12 alloy (a), ( a') and annealed for 1 h at
given temperatures (b) - (f) and (b') - (n .

not contain these spectral components.


In all CEMS and transmission spectra of the Fesl Zr7 8 12 alloy annealed at
T A = 500. 550 and 600 'c a spectral component with Hhf ~ 27. 5 - 29 T and
e5~0. 04 mm/s is observed. This spectral component resembles the one
corresponding to the interfacial regions in the Fe/Zr multilayers studied
recently (Kopcewicz et al.. 1994b). Hence. we attribute this sextet to the
grain boundary regions between the nanocrystalline bee-Fe and the amorphous
host.
Higher annealing temperature (T A = 780 'C) causes complete bulk
crystallization of the FeSl Zr78 12 alloy. The amorphous phase disappears and
the spectra. both transmission and CEMS. consist only of one sextet
characteristic of the a-Fe phase (Fig. 5. 31f. f ') .
The bulk and surface crystallization of the Fe79 Zr7 8 12 CU2 alloy proceed in
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 235

a distinctly different way. As before, the as-quenched Fe79 Zr7 B 12 CU2 alloy is
completely amorphous (Fig. 5. 32a); there are no traces of crystallization at
Transmission CEMS
1.00

0.95

1.00

0.98

"0
'03 1.00 "
0
'03
.'" (c) .'"
'" '"
" "
<I,) <I,)
> 0.98 >
.~ .~
Q) Q)
~ 1.00 ~

(d)
1.01

0.95
1.00
1.00
(e) (e')
1.02

0.96 600'C 1.00


--{) -3 0 3 6 --{) -3 0 3 6
Velocity (mm/s) Velocity (mm/s)

Figure 5.32 Mossbauer transmission (a) - (e) and corresponding CEMS (a') - (e')
spectra recorded for the as-quenched Fe'9Zr,B'2Cu2 alloy (a), (a') and annealed for 1 h
at given temperatures (b) - (e) and (b') - (e') .

the surface (Fig. 5. 32a'). Annealing at T A =


430 'c similarly to the FeZrB
alloy causes onset of the surface crystallization and the bcc-Fe phase appears
(Fig. 5. 32b'). At T A = 500 'c bulk crystallization is already induced resulting
in a formation of the bcc-Fe phase (Fig. 5. 32c). Surface crystall ization is
much more pronounced than the bulk one as shown by the CEMS spectrum in
Fig. 5. 32c'. However, in distinction to the FeZrB alloy the paramagnetic OS
doublet observed at the surface in the Fe81 Zr7 B 12 sample (Fig. 5. 31 c', d') is
not detected in the Fe79Zr7B,2Cu2 alloy (Fig. 5. 32c'). After annealing of the
Fe79 Zr7 B 12 CU2 alloy at T A =
550 'c, in addition to the bcc-Fe phase, a
paramagnetic phase is formed. The single-line spectral component with
236 Michat Kopcewicz

{j = - 0.06 mm/s detected both in transmission (Fig. 5. 32d) and CEMS


(Fig. 5. 32d') spectra allowed the identification of this phase as y-Fe. This
phase is clearly seen also after annealing at 600 'c (Fig. 5.32e, e'). The y-
Fe phase was not observed in the Feel Zr7 B 12 alloy at any annealing
temperature (Fig. 5.31). On the other hand the crystalline Fe2B phase,
clearly seen in the CEMS spectrum of the Feel Zr7 B 12 alloy annealed at T A =
600 'c (Fig. 5. 31 e'), was not observed for the Fe79 Zr7 B 12 CU2 alloy at any
annealing temperature (Fig. 5.32). The interfacial component is considerably
more abundant in the Cu-containing alloy annealed at T A = 430 - 600 'c as
compared to the Feel Zr7 B 12 one.
Annealing at higher temperature ( T A = 780 'C) causes complete
crystallization of the FeZrBCu alloy. The transmission and CEMS spectra
consist of ex-Fe sextet only, similarly to the case of the FeZrB alloy.
Relative content of the bcc-Fe phase is considerably larger in the
FeZrBCu alloy than in the FeZrB one as revealed both by the transmission and
CEMS spectra (compare Figs. 5.31a-e and 5.32a-e, and Figs. 5.31a'-
e' and 5. 32a' - e', respectively).
The results are summarized in Figs. 5. 33 and 5.34 for the Feel Zr7 B 12 and
Fe79 Zr7 B 12 CU2 alloys, respectively, which show the relative abundance of
various Fe-containing phases detected in the transmission (Figs. 5. 33a,
5.34a) and CEMS (Figs. 5.33b, 5.34b) spectra. The residual amorphous
phase is, at all anneal ing temperatures, much more abundant in the bulk than
at the surface. The CEMS results for the Feel Zr7 B 12 alloy are collected for the
shiny side (Fig. 5.33b) and the dull side (Fig. 5. 33c). As can be seen, on
the dull side the same phases appear at the same temperatures as on the shiny
side, only their relative abundance is lower. The spectral contribution of the
interfacial grain boundary component is considerably more abundant in the
Fe79 Zr7 B 12 CU2 alloy annealed at T A = 430 - 600 'c than in the case of
Feel Zr7 B 12 alloy. Its spectral contribution increases with increasing T A from
4 % to about 15 % at T A = 500 and 600 'c , respectively (Fig. 5.34), while for
the Feel Zr7 B 12 alloy it does not exceed 5 % (not included in Fig. 5. 33). The
content of this phase is somewhat larger at the surface than in the bulk. The
spectral contribution of the grain boundary component is directly related to the
degree of crystallization and increases with the increasing number of
nanograins formed in the amorphous host. The onset of crystall ization at the
sample surface, occurring in both alloys at lower annealing temperatures (T A )
than the bulk crystallization temperatures (T Xl ) , can be interpreted in a
similar way as in earlier studies of the surface crystallization of conventional
amorphous alloys (Gonser et aI., 1983; Ok and Morrish, 1981). Most
probably some boron atoms escape from the surface region during melt
spinning process and during annealing. The decrease of boron content in the
amorphous Feel Zr7 B 12 and Fe79 Zr7 B 12 CU2 alloys decreases crystallization
temperature as revealed by the DSC and Mossbauer measurements for
Fe93-x-y Zr7BxCuy alloys (Fig. 5. 1). Hence, the crystalline phases,
especially bcc-Fe, appear at the sample surface at lower T A than in the bulk.
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 237

As-quenched Transmission
100 7-

~
80
~ v Amorph.
"c:
u
60 0 bcc-Fe
'c:"
"0
:::J
<> Fe 2B
.D 40 l> QS doublet
"'>"
.~
20
vc:
0
0 400 500 600 700 800
Annealing temperature (OC )
(a)

As-quenched CEMS
100 V'- Shiny side
~

e 80
"uc:
'"c:
"0
60
:::J
.D
40
">'"
.~

v 20
c:
0
0 400 500 600 700 800
Annealing temperature (OC )
(b)

As-quenched CEMS
100 V'- Dull side
~

~
~ 80
"uc:
'"c: 60
"0
:::J
.D

"'">
40
.~

vc: 20
0
0 400 500 600 700 800
Annealing temperature (OC )
(c)

Figure 5. 33 Relative abundance of various phases determined for Fesl Zr7 B '2 alloy (a) by
transmission, (b) by CEMS (shiny side) and (c) by CEMS (dull side) techniques.
238 Michat Kopcewicz

As-quenc he d Transmission
100 v-
,....,
~ 80
"uc: 'V Amorph.
'"c:
"0
::l
60 0
t;
bee-Fe
Y-Fe
.D
gr. bound.
"'"
40 0
;>
.~
20
"
0:::

0
0 400 500 600 700 800
Annealing temperature ('C )
(a)

As-quenched CEMS
100 <;>- Shiny side
,....,
:? 80
~

"
u

"0 60
c:
::l
.D
40
"'"
;>
.~
20
"0:::

0
0 400 500 600 700 800
Annealing temperature ('C )
(b)

Figure 5. 34 Relative abundance of various phases determined for Fe79 Zr7 8 ,2 CU2 alloy
(a) by transmission, (b) by CEMS (shiny side of the ribbon) techniques.

Kinetics of the surface and bulk crystallization were studied for the
FesIZr7Bl2 alloy by means of the combined CEMS and y-transmission
Mossbauer technique. Amorphous alloy was annealed at 600 C for times
ranging from 15 s to 120 s and for 1 h (Grabias and Kopcewicz, 1998). The
transmission and CEMS spectra recorded for the Fesl Zr7 B '2 alloy annealed as
a function of anneal ing time at T A = 600C are shown in Figs. 5.35 and 5.36,
respectively, together with the P (H) distributions extracted from the
corresponding spectra. The transmission spectra (Fig. 5. 35a - f) reveal the
formation of the bee-Fe phase in the bulk of the sample already after 15 s of
annealing (Fig. 5.35a) as evidenced by the appearance of sharp lines. The
relative contribution of this sextet is slightly larger for 1 h of annealing
(Fig. 5.350 than after annealing for 60 - 120 s (Fig. 5. 35c - e). The P (H)
distributions (Fig. 5. 35a' - f () consist of a peak at about 14 T corresponding
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 239

0.96 0
1.00

0.95
0
1.00

t: 5
0 0.98
'Vi
if>
.~
0
~
Q)
1.00 :
Cl..,
>
.~ 5
Q)
~
0.96
0
1.00

---{) 0 6 0 10 20 30 40
Velocity (mm/s) H(T)

Figure 5. 35 Transmission M6ssbauer spectra (a) - (f) and the corresponding P ( H)


distributions (a') - (n for the Fesl Zr7 8 12 alloy annealed at 600"C for 15 - 120 s and for 1 h.

to the retained amorphous phase, and the second smaller peak at higher
hyperfine fields (27 - 29 T) which is attributed to the interfacial regions
between the bcc-Fe grains and the amorphous matrix in the same way as
discussed in Section 5. 5. 1. The relative atomic fractions of Fe as the
amorphous, bcc-Fe and interfacial phases calculated from the transmission
spectra presented in Fig. 5. 35 are shown in Fig. 5. 37a. The amorphous
phase strongly dominates in the bulk and its spectral contribution saturates at
about 80 % for the sample annealed for 60 s. The relative content of the bcc-
Fe phase increases with annealing time and saturates at about 8 % after 60 s
anneal. The CEMS spectra recorded for all annealed samples (Fig. 5. 36a - f)
clearly reveal much stronger surface than bulk crystallization (Fig. 5. 35a-f).
The spectral component related to the amorphous matrix strongly decreases
with the annealing time (Fig. 5. 37b, c). The relative content of the bcc-Fe
240 Michat Kopcewicz

15 s
1.01
5

1.00
1.01

1.00
1.02
5
0
.;;;
c
B 1.00
.::
Q)
>
.~

Q; 1.01
IX:

1.00

1.01

1.00
1.01

1.00~.~~~~
~ 0 6 o 10 20 30 40
Velocity (mm/s) H(T)

Figure 5.36 CEMS spectra (a) - (f) and the corresponding P (H) distributions (a') -
(n for the FeS,Zr7B,2 alloy annealed at 600'C for 15- 120 s and for 1 h.

phase increases dramatically for very short of annealing( 15 and 30 s) and then
almost saturates at about 40 % after 60 s anneal (Fig. 5. 37b - c). However,
other phases, not observed in any transmission spectrum for bulk
crystallization, appear at the sample's surfaces. In the CEMS spectra in
Fig. 5. 36a - c a significant paramagnetic spectral component (a quadrupole
doublet with the quadrupole spl itting OS = 0.80 mm/s and IS = O. 34 mm/s) is
observed. It contributes to 28% (shiny side, Fig. 5.37b) or 16% (dull side,
Fig. 5.37c) for the sample annealed for 15 s. The spectral contribution of this
doublet decreases as annealing time increases and it disappears after 120 s
and longer (1 h) anneals (Figs. 5. 36e, f and 5. 37b, c). The origin of this
nonmagnetic spectral component is related, as in the case of FeZrB ( Cu)
alloys (Fig. 5. 31 ), to the iron-poor paramagnetic amorphous Fe-B phase
which decomposes at longer annealing times.
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 241

100~Trans.
bulk
80

8 60

.g""
t:
40
<l.l
>
.<ij 20
Qj

O~
0:::

o 40 80 120 3600
Annealing time (s)
(a)
,<
100 CEMS
Shiny side
~
80
C
8t: 60
"""t:
E 40
">
<l.l

.<ij 20
Qj
0:::
o
o 40 80 120 3600
Annealing time (s)
(b)

'V Amorph.
o bee-Fe
~ <> Fe2B
~ o Interf.
8t: '" QS
"t:
""E 40
"
<l.l
>
.<ij 20
Qj
0:::
o
o 40 80 120 3600
Annealing time (s)
(e)

Figure 5.37 Kinetics of crystallization: relative abundance of various phases determined


for FeS,Zr7B,2 alloy (a) by transmission, (b) by CEMS (for shiny side) and (c) by CEMS
(for dull side) techniques.
242 Michat Kopcewicz

The amorphous phase is still retained in the sample annealed for 120 s. Its
spectral contribution decreases with anneal ing time to less than 40 %
(Fig. 5.37b, c). Two independent P(H) distributions (Fig. 5.36a'-f')
were extracted from the broadened sextet in the CEMS spectra. The P ( H)
ranging from 0 to about 20 T corresponds to the retained amorphous phase.
The second P ( H) ranging from 20 T to about 35 T, is attributed to the
interfacial regions between the bcc-Fe grains and the amorphous matrix. The
relative content of the Fe atoms in the interfacial regions saturates at about
17 % already after 30 s anneal and remains the same for longer anneal ing times
(Fig. 5. 37b, c). The change in the structure of the retained amorphous phase
is clearly seen as the splitting of the single peak in the P ( H) distribution
obtained for 15 s of annealing (Fig. 5. 36a') into several peaks when the
abundance of the bcc-Fe phase increases (Fig. 5. 36d' - f '). This suggests
strong composition fluctuations in the amorphous matrix due to progressing
crystall ization.
After long annealing time (1 h) the crystalline Fe2 B phase is formed at the
sample's surfaces in addition to the phases is discussed above. The
characteristic spectral components of the tetragonal Fe2 B phase, with two
characteristic sextets (Fig. 5.360, contribute to about 40 % and 30 % for the
shiny and dull sides of the sample, respectively. Spectral contribution related
to this phase increases with annealing time (Fig. 5. 37b, c), and as it was not
observed in the transmission spectra, it turns out that Fe2 B was formed at the
surfaces of the sample only.
The qualitative changes observed in the transmission and CEMS spectra
as a function of annealing time at 600 'c are similar to those recently observed
as a function of annealing temperature for 1 h anneals (Fig. 5.31a'-n. This
means that the formation of a given phase is related to the degree of
crystallization rather than to the annealing temperature alone.
The comparison of the CEMS spectra for the shiny and dull sides of the
samples reveals that the amorphous matrix is less abundant at the shiny side.
Much stronger crystallization at the surfaces as compared with the bulk is most
probably related to the loss of some boron from the surface regions during the
melt spinning process and during annealing that leads to the decrease of the
crystall ization temperature of the FeZrB alloys, as revealed by the DSC
(Fig. 5.1) and Mossbauer experiments (see Section 5.5.1).
The CEMS measurements were performed recently for various
NANOPERM-type alloys of the composition Feso M7B 12 CUI (M: Ti, Nb, Ta)
(Grabias et al., 1999a). For Feso Tb B 12 CUI and Feso Nb 7B I2 CUI alloys the
differences between the surface and bulk crystallization were smaller than
those observed for Fe79 Zr7 B 12 CU2 alloy discussed above. However, also in Ti-
and Nb-containing alloys the spectral contributions of the bcc-Fe nanograins
were significantly larger in the CEMS spectra than in the corresponding
transmission spectra. This behavior is preserved throughout the entire
annealing temperature ( T A) range. The evolution of the retained amorphous
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 243

phase with annealing temperature is clearly seen both in CEMS and in


transmission spectra. The average hyperfine field characteristic of the
amorphous phase increases with increasing T A and clearly exceeds that of the
as-quenched FeTiBCu and FeNbBCu alloys. This strongly suggests that the
change of the composition of the retained amorphous phase (due to the
formation of the bee-Fe grains) towards the iron-poorer one leads to the
increase of the hyperfine field at room temperature which is most probably
related to the increase of the Curie temperature of the retained amorphous
phase with the decrease of iron content in that phase. From the CEMS and
transmission spectra the relative atomic fraction of Fe as bee-phase was
determined from the spectral area of the component with the hyperfine field of
about 33 T characteristic of the bee-Fe phase. The results obtained for all
alloys studied vs. annealing temperature are shown in Fig. 5.38. The relative
spectral contribution of the bee-Fe sextet increases with increasing annealing
temperature at the expense of the P ( H) component related to the retained
amorphous phase, indicating the increase of the volume fraction of the bee-Fe
phase. In all cases the crystall ine phase is more abundant at the surface than
in the bulk of the sample at each anneal ing temperature (F ig. 5. 38) .
V> V>

E "V> 70
.8 0 Surface E "V> 70
.8
'" '" 60
"..<::
u..Q. 0 bulk '""..<::'" 60
u..Q.

~
'- " 50 '- " 50
ol.r;< ol.r;<
c u 40 c u 40
o u o u
';: ~ 30 ';:; .J::J 30
u " 20 M=Ti ~~
"'..<::
<1::- <1::- 20
--: ~ 10 -=~ 10
"
~
o 400 ~ " 0
450 500 550 600 400 450 500 550 600
Annealing temperature ('C ) Annealing temperature ('C )

V> V>
E E
o "lfj 70
., o "
~ ~
70
60 ".e 60

~ ~
"..<::
u..Q. u..Q.
'- " 50
ol.r;< '-
ol.r;<" 50
c u 40 C u 40
o u
.;:;..0 30
o u
'';::'.J::J 30
"'.e 20
U " u "
"'.e 20
<1::- <1::-
....: lfj 10 -= ~ 10
~ " 0
400 450 500 550 600
~" 0
400 450 500 550 600
Annealing temperature ('C ) Annealing temperature ('C )

Figure 5.38 The relative abundance of the bee-Fe phase determined for FeaoM7B'2Cu,
alloys.

Recently the first CEMS measurements were performed for HITPERM


alloy with the composition Fe44 C0 44 Zr7 B4Cu, (Kopcewicz et al., 2001). The
CEMS spectra recorded for the amorphous alloy revealed the traces of
crystallization on the shiny side of the as-quenched ribbon while the dull side
and the bulk were completely amorphous. The CEMS spectra measured for the
244 Michat Kopcewicz

sample annealed at 650 'c for 1 h have shown a greater extent of


crystall ization on surfaces than in the bulk of the sample. Both surfaces of the
annealed ribbon show similar characteristics. The surfaces are completely
crystallized to the bcc-FeCo phase without the residual amorphous phase,
which was clearly present in the bulk of the sample, as revealed by the y-
transmission M6ssbauer spectra.
The CEMS results discussed above provide detailed information regarding
the surface crystallization of amorphous precursors and show the power of the
CEMS technique in such studies.

5.7 Short Range Order in Amorphous Precursors-


the rf-Mossbauer Study

The problem of the short range order (SRO) in amorphous alloys still remains
unsolved despite extensive experimental and theoretical efforts. The
techniques most commonly used in the study of local structure were X-ray and
neutron diffraction. Very powerful techniques used for the determination of the
structural ordering in amorphous alloys are based on hyperfine interactions. A
prominent example of such methods is the M6ssbauer spectroscopy. The most
direct probe of the local structure is provided by the electric quadrupole
hyperfine interaction which is determined by the nearest environment of the
M6ssbauer nuclei. However, most amorphous iron-based metal-metalloid
alloys are ferromagnetic at room temperature and the magnetic dipole
hyperfine interaction is usually much stronger than the electric quadrupole one,
so the quadrupole interaction cannot be studied directly. The coexistence of
the magnetic dipole and electric quadrupole interactions leads to complex
M6ssbauer spectra, so the conclusions regarding the SRO, based usually on
the magnetic hyperfine field distributions, are not always reliable. To obtain
information about the local structure from the electric quadrupole interaction it
is best to separate this interaction from the magnetic one. As it was discussed
in Section 5.2.3, this can be done by an unconventional rf-M6ssbauer
technique which makes use of the rf collapse effect induced by the radio-
frequency (rt) magnetic fields in the material investigated. The collapse of the
magnetic hyperfine splitting observed in the rf-M6ssbauer spectra causes it to
happen that in place of the complex magnetically split six-line pattern the
quadrupole split doublet (OS) appears though the sample is in the
ferromagnetic state. Therefore the distributions of the quadrupole splittings
and the information regarding changes of the local structure can be readily and
reliably obtained. The application of the rf-M6ssbauer technique for the study
of the short range order (SRO) in conventional amorphous alloys is extesively
reviewed by Kopcewicz (1987, 1991).
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 245

Here we will discuss the application of the rf-M6ssbauer technique in the


investigation of the structural changes in amorphous Fe93-x-yZr7BxCuy(x=4,
6,8,12; y=O, 2) (Kopcewiczetal., 1998a) and FesoM7B12CU1(M: Ti, Ta,
Nb, Mo) (Kopcewicz et aI., 1999a) alloys. These alloys were studied
extensively because they serve as amorphous precursors for the formation of
NANOPERM-type nanocrystalline alloys (see Section 5.5.1).
The SRO was studied for the amorphous Fe93- x- yZr7 B xCUy (x = 4, 6, 8,
12; y=O, 2) alloys as a function of alloy composition via quadrupole splitting
distributions, P ( OS ), derived from the rf collapsed M6ssbauer spectra
(Kopcewicz et al., 1998a). The amorphous structure of the as-quenched
ribbons, prepared by the melt spinning technique, was checked by X-ray
diffraction measurements. The rf-M6ssbauer measurements were performed
for the samples exposed to the rf field of 20 Oe at 60. 9 MHz, sufficient to
average to zero the magnetic hyperfine field and to induce the complete rf
collapse effect. The P ( OS) distributions were extracted from the rf collapsed
spectra by using the histogram method (Hesse and RObartsch, 1974; LeCaer
and Dubois, 1979) in the same way as in the earlier studies (Kopcewicz
et aI., 1986a, 1986b, 1993, 1995d).
All amorphous FeZrB( Cu) alloys investigated, except for FeS9 Zr7 B4 , are
ferromagnetic at room temperature. Since the rf collapse effect occurs only in
the ferromagnetic materials, the FeS9 Zr7 B4 alloy is not affected by the rf field
so the M6ssbauer spectra recorded without the rf field and during rf exposure
are almost identical. The slight differences in the center shift of the quadrupole
doublet are related to some rf heating of the rf exposed sample. The P ( OS)
distributions extracted from the spectra recorded with and without the rf field
are identical.
The rf collapsed spectra obtained for various alloy compositions are
shown in Fig. 5. 39a - g. They consist of an asymmetric broadened quadrupole
doublet. The asymmetry was well reproduced by introducing in the fitting
procedure a linear correlation between the isomer shift and the quadrupole
spl itting. The shape of the spectra changes considerably; the OS doublet
becomes less resolved for higher boron content. This change is clearly
reflected in the P ( OS) distributions which consist of two well defined peaks:
one at low OS values with the corresponding average quadrupole splitting of
about O. 12 mm/s and the second, a much broader peak at about O. 55 mm/s
(Fig. 5. 39a' - g'). The relative contributions of these peaks to the total
P ( OS) distribution vary with boron content. The peak at low OS values
markedly increases with decreasing boron content at the expense of the peak
at high OS. The peak at high OS becomes narrower at low boron content and
shifts towards smaller OS values (compare Fig. 5. 39a', g'). However, the
presence of Cu does not affect the shape of P ( OS ), regardless of boron
content (Fig. 5. 39a' -f ').
The shape of P ( OS) distributions observed for all alloy compositions
strongly suggests that the structure of the amorphous FeZrB ( Cu) alloys is
246 Michat Kopcewicz

150 .
x=12 100 .....: ... (a')
y=O
50
.................
0 ..........
100
.....
x=12
(b')
y=2
50
......
0 ..........
x=8 100
" .. ".
(c')
y=O
50
" .................
0
'Vi
0
'. .. ....
'6'" x=8 100 (d')
'" G
g" y=2 g 50
'>.,
">
.~ o ...............
Q) 150
e<: x=6 (e')
y=0 100
50
o ................
150 ........
o,'

100 (f')
50
o ...... .................
x=4 ....
100 (g')
y=O
50
o'------'--------'.:.::.::...,...-"-
-I 0 I 2 o 0.5 1.0 1.5
Velocity (mm/s) QS(mm/s)

Figure 5. 39 The rf-collapsed Mossbauer spectra of the amorphous Fe93-x-y Zr7 B x CUy
(x = 12,8,6; y = 0,2) alloys (a) - (g) and the quadrupole splitting distributions (a') -
(g') extracted from the fits of the corresponding spectra (a) - (g) .

inhomogeneous. There are two preferential SRO: one in which Fe atoms


experience a more symmetrical local atomic arrangement (small OS values)
and the other with a more distorted local structure (high OS values). The
relative content of iron atoms contained in the more symmetric SRO increases
with increasing Fe abundance (decreasing boron content) in the FeZrB (Cu)
alloys. Higher boron content induces stronger distortion of the structure which
is most probably related to the strong repulsive interaction between metalloid
atoms. The clearly different OS values strongly suggest composition
fluctuations in the amorphous phase. Small OS values correspond to iron-rich
FeZrB regions. The decrease of the local iron content leads to the increase of
OS. This trend in OS vs. iron concentration in the amorphous phase was
clearly observed in the earlier systematic studies of the rf collapse effect in
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 247

various iron-based amorphous alloys (Kopcewicz et aI., 1986a, 1986b;


Kopcewicz, 1991).
Similar inhomogeneities of the structure in amorphous FeZrB (Cu) alloys
with 6 % B have been suggested recently on the basis of the magnetic
hyperfine field distributions obtained from the conventional Mossbauer spectra
measured at low temperature (Duhaj et aI., 1996). However, the SRO and its
changes with alloy compositions are considerably more pronounced in our rf-
Mossbauer studies than in the conventional Mossbauer experiments.
The rf-Mossbauer results for amorphous Fe-M-B-Cu alloys, in which as M
the Ti, Ta, Nb and Mo atoms were introduced, are shown in Fig. 5. 40

1.0
(a) M=Ti (a')
JOO
0.9
0
1.0
(b) M=Ta (b')
JOO
c:
0
.;;;
'" 0.9
.~ 0
g 1.0 GJ
Q)
(c)
g
> M=Nb (c')
.,
.~

et: 0.9
"'- JOO

0
1.0

(d) M=Mo 100 (d')

0.8
0

-2 -J 0 J 2 0 0.5 1.0 1.5


Velocity (mm/s) QS(mm/s)

Figure 5.40 The rf-collapsed M6ssbauer spectra of the amorphous FesD M7 8'2 Cu, alloys
(a) - (d) and the corresponding quadrupole splitting distributions (a') - (d').

(Kopcewicz and Grabias, 1996). In distinction to the FeZrBCu case, the


p ( OS) distributions are almost identical for all alloy compositions, indicating
that such substitutions do not significantly affect the short range order in Fe-M-
B-Cu alloys. This result is not surprising because the atomic radii of the
alloying elements are very similar: 1.47, 1.49, 1.46 and 1.39 A for Ti, Ta,
Nb and Mo, respectively. Similarly to the FeZrBCu alloys, the amorphous
structure of Fe-M-B-Cu alloys is inhomogeneous as revealed by the P ( OS)
distributions consisting of two peaks (Fig. 5. 40), which suggests the
presence of iron-rich and iron-poor regions.
In addition to the rf-Mossbauer studies of the amorphous alloys, the
248 Michat Kopcewicz

conventional M6ssbauer measurements (without rf field) were performed


(Kopcewicz and Grabias, 1996). The room temperature measurements
performed for both series of alloys reveal that the hyperfine fields vary
significantly with alloy compositions. Most of the alloys are ferromagnetic at
room temperature, except for Fes9 Zr7 B 4 and FeSD M07B 12 CUI' Figure 5. 41
shows the spectra recorded at room temperature for the FeZrB alloys with
various boron content. The changes of the hyperfine fields reflect the
differences in the Curie temperatures (T c) of the alloys studied, and are not
related to the SRO. In order to extract information regarding the local structure
from the conventional M6ssbauer measurements data have to be collected at
temperatures much lower than T c of all alloys studied to avoid the influence of
T c on the hyperfine field distributions. Therefore the M6ssbauer measurements
were performed at 78 K (Fig. 5. 42). As can be seen from Fig. 5. 42, all
spectra consist of a broadened sextet characteristic for amorphous alloys. The
1.00

(a) x=4

0.80
1.00

"0
'iii (b')
'" 0.95
.~

" 1.00
<l)
>
.~

" 0.95
~
(c')

0.95
--{j 0 6 o 40
Velocity (mm/s)

Figure 5. 41 M6ssbauer spectra recorded at room temperature for the amorphous


Fe93-xZr7Bx alloys (a) - Cd) and the corresponding hyperfine field distributions Cb') -
Cd') .

hyperfine field distributions, P (H), extracted using the same method as for
the P ( OS) distributions (Hesse and Rubartsch, 1974; LeCaer and Dubois,
1979) consist of two well defined peaks: the main one at about 25 T and the
smaller one at low hyperfine fields (about 12 T). The positions of both peaks
in P (H) does not change with alloy compositions (Fig. 5. 42). The relative
contribution of these peaks changes with boron content; the peak at 12 T
decreases slightly as boron content increases, which may suggest a change in
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 249

the SRO as the alloy composition changes. However, the changes in the
PC H) distributions are less distinct and less credible than those observed in
PC OS).
1.00

0.95

1.00 o
<::
.~ 5
.
;g 0.95
.s
(\)
1.00
>
.~

~
0.95
1.00 o

0.95

-606 0
Velocity (mm/s)

Figure 5.42 M6ssbauer spectra recorded at T = 78 K for the amorphous Fe93- x Zr, Bx
alloys (a) - (d) and the corresponding hyperfine field distributions (a') - (d').

The low temperature Mossbauer measurements performed for Fe-M-B-Cu


alloys show that the PC H) distributions are very similar regardless of M atom
substitution. This finding is in agreement with the rf-Mossbauer results
discussed above, and confirms that Ti, Ta, Nb and Mo substitutions do not
significantly affect the short range order in the Feso M7 B 12 CUI amorphous alloys.
The rf-Mossbauer results are much more reliable for the determination of
the changes in the short range order vs. alloy composition because the electric
quadrupole hyperfine interaction is directly determined by the local structure,
while the magnetic hyperfine field is only indirectly related to the symmetry
experienced by Fe atoms, and may be significantly affected by the relation
between the temperature of the measurement and the Curie point for a given
material. The rf-Mossbauer technique provides a more accurate and sensitive
probe of the SRO than conventional X-ray and neutron diffraction methods.
250 Michat Kopcewicz

5. 8 Determination of the Grain Size by XRD and SAXS


Techniques-Correlation with the Mossbauer Results

The X-ray diffraction (XRD) technique, commonly used for characterization of


the crystallization process of amorphous precursors allows the estimation of the
size, L, of nanograins, using a well known Scherrer formula (Klug and
Alexander, 1974). In order to estimate credibly the grain size from the
broadening of the XRD line, the broadening should result from the refinement of the
grains and not from other phenomena, e. g., strain in the material. The limitation of
this technique is also related to the fact that grain size determined using the Scherrer
formula concerns the spherical grains, and the distortion of the shape of the grains
introduces uncertainty of the values obtained in this way.
As an example, determination of the grain size in the nanocrystalline
Fe~l3-x-yZr7BxCuy (x = 6, 8, 12; y = 0, 2) (Kopcewicz et aI., 1997a) is
discussed. These amorphous precursors were annealed for 1 h at temperatures
ranging from 430 to 780 C , so the growth of the nanocrystalline phase was fairly
slow, and it can be expected that possible stresses related to the formation of the
crystalline grains have been relaxed. The Mossbauer results obtained for these
alloys were discussed in Section 5. 5.
The X-ray diffraction measurements were performed for all alloys in the
as-quenched state and after annealing at 500 and 550 C. All as-quenched
alloys were completely amorphous as shown by the corresponding X-ray
diffraction patterns consisting of broad peaks at 28;;::,;: 44 and about 78.
Annealing at T A ~ 500C results in the formation of the bee-Fe phase in all
alloys. Data from several samples are shown in Fig. 5.43. Distinct diffraction
peaks appear in the XRD patterns whose positions agree well with the bee-Fe
phase. The intensity of these peaks increases with increasing T A showing the
increase of the volume fraction of the bee-Fe phase at higher T A. From the
XRD patterns the average size of the bee-Fe grains, L, was estimated using
the Scherrer formula with a Scherrer constant K = 0.9. The linewidths of the
( 110), (200), (211) and (220) peaks in the 500C and 550 c patterns were
used to find an average value, <L >. The results are collected in Table 5. 3
(Kopcewicz et aI., 1997a).
As can be seen from Table 5.3, addition of 2 % Cu to the alloy with fixed
boron content leads to smaller grains for all values of x, particularly for x =
12. Increase of the anneal ing temperature causes the growth of the grains for
x = 12 but little or no change for the other samples. Since the <L > value for
the FeBI Zr7 B 12 alloy annealed at 500C is determined from a very weak bee-Fe
peak, its experimental error is considerably larger than in all other cases: this
alloy is almost completely amorphous (F ig. 5. 43, Table 5. 3). However, in
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 251

80

#4-550C

60

S
""'
~
0
'v;
t:
<I.l
40
.s
~
~

20

O'--_....L..-_--'-_-'-_ _'-_....L..-_--'-_----'
30 40 50 60 70 80 90 100
2ee )
Figure 5. 43 The X-ray diffraction spectra for sample # 1 (x = 12, Y = 0), # 2 (x = 12,
Y = 2) and # 4 (x = 8, y = 2) annealed at 500 and 550 e (see table 5.3).

Table 5.3 The bcc grain sizes determined by XRD and SAXS: (L) - the average grain size
determined from XRD patterns using the Scherrer formula; (D) - the average grain size
determined by SAXS; Q - the integrated SAXS intensity, proportional to the volume fraction
of nanoparticles; A I q3 describes the Porod tail.

Sample Sample composition XRD results SAXS results


TAce)
number. (L)(nm)
x y Q( 10 23 eu/cm 3 ) A (O)(nm)

500 12 19 2.9 5
1 12 0 550 14 2.8 3.8 7
600 21 3.2 4.0 10
500 5 7.1 2.0 4
2 12 2 550 7 9.8 2.0 4
600 9 13.8 1.2 3
252 Michat Kopcewicz

Table 5. 3 Continued from page 251


Sample Sample composition XRD results SAXS results
TACC)
number x <L> (nm)
y Q (10 23 eu/cm 3 ) A <O)(nm)

500 11 5.4 30 5
3 8 0
550 11 8.7 1.5 5
500 8 7.1 1.3 4
4 8 2
550 9 10.8 2.0 4
500 13 30 4.2 6
5 6 0
550 12 5.4 2.0 6
500 11 6.5 3.0 4
6 6 2
550 11 7.9 6.0 5

all FeZrB(Cu) alloys studied, the grain sizes ranged from about 5 to 20 nm
depending on the alloy composition and annealing conditions; thus the bee-Fe
phase is in all cases nanocrystall ine.
The grain size was also determined in the same way in the Feso M7 B 12 CUI
(M: Ti, Ta, Nb, Mo) alloys (Kopcewicz et al., 2001). The grain size, L,
was determined from the broadening of the XRD peak at 2 e """
44 related to
the bee-Fe. The size of the grains in Ti-containing alloy depends on the
annealing temperature. At low annealing temperatures, T A ~ 470 'c, the
grains are fairly large; L """8.6 nm. At higher annealing temperatures smaller
grains are formed. The grain size decreases from 8. 1 to 7.8 nm for T A = 520
and 570 'C, respectively. Further increase of T A causes the growth of the
grains (L = 8.8 nm for T A = 620 'c ). The XRD results agree very well with the
rf-M6ssbauer observations (Fig. 5. 29) and with the measurements of
coercivity vs. T A (Fig. 5. 17). The smallest bee-Fe grains were formed in Zr-
and Nb-containing alloys. The average size of the grains is about 7 nm and
9 nm for FeZrBCu alloy annealed at 550 'c and 600 'c , respectively, and 5.9
and 6. 2 nm for the FeNbBCu alloy annealed at 520 and 570 'c , respectively.
The rf-M6ssbauer measurements revealed that these alloys have the best soft
magnetic properties (smallest anisotropy) (Fig. 5. 30). Considerably larger
grains were formed in Ti-containing alloy, in agreement with the rf-M6ssbauer
data. However, relatively large grains were formed in the Mo-containing alloy
(L """9.2 and 8.6 nm for T A = 520 and 570 'c, respectively) for which fairly
small magnetic anisotropy was found.
Another method that can provide microstructural information and allows
the calculation of the grain size is the small-angle X-ray scattering (SAXS).
The SAXS technique has not been commonly used for characterization of the
metallic amorphous (Lamparter and Steeb, 1988; Kamiyama et aI., 1992)
and nanocrystalline alloys (Kopcewicz et aI., 1996). Small angle neutron
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 253

scattering (SANS) was recently applied to FeCuNbSiB nanocrystalline alloys


(Kohlbrecher et al., 1995; Ohuma et aI., 1995) and an Feso B20 alloy (Deriu
et aI., 1996). The SANS technique was combined with TEM and allowed the
evaluation of the size of the bee-Fe nanocrystals within larger Fe3 B crystall ites
formed as a result of crystallization of the amorphous Feso B20 alloy.
SAXS measurements were performed for the Fe93-x-yZr7BxCuy(x=6, 8,
12; y=O, 2) alloy (Kopcewicz et al., 1996, 1997a) on the same specimens
used for the Mossbauer study (see Section 5.5.1 and 5.5.2). They were thin
enough (-15-20 IJm) for this transmission method as well. A Kratky-type,
line-focus, small angle system (Glatter and Kratky, 1982), attached to a Cu
rotating anode X-ray generator was used to collect scattered intensity data
over a momentum transfer range from q = O. 1 to 6 nm -1, where q = (411"/ A )
sine, 0=0.15418 nm, and 2e is the scattering angle). This range allows
detection of electron density fluctuations on a size scale from about 30 to 1 nm.
Details of the apparatus and data treatment are described elsewhere
(Williamson et aI., 1989; Williamson, 1995). One feature of the present Fe~
rich samples that was carefully taken into account was the large amount of Fe
fluorescence caused by the Cu X-rays. The energy window of the Cu-K" X-ray
(8.05 keV) counting channel was shifted on the low energy side to reduce the
intensity of Fe X-rays (6. 4 keV) to an acceptable level. A constant SAXS
intensity was observed at the highest angles due in part to the residual Fe
fluorescence and in part to diffuse scattering mechanisms from amorphous
alloys (Wi II iamson, 1995). Such a constant intensity is readi Iy subtracted to
obtain the SAXS from microstructural features of interest such as medium range
order (Lamparter and Steeb, 1988), crystallites (Kamiyama et aI., 1992) or
microvoids (Williamson, 1995). All intensities were converted into absolute
intensity units (electron units = electrons/atom = e/ a) based on a calibration
of the SAXS system using the X-ray scattering from H20 (Hendricks et al.,
1974) .
Let us discuss first the results obtained for Fesl Zr7 B 12 and Fe79 Zr7 B I2 CU2
alloys (Kopcewicz et aI., 1996). Figure 5.44 compares the SAXS intensities
of the FeZrB and FeZrBCu samples in the as-quenched states and after
identical anneals. The constant background intensity due to the Fe
fluorescence and other diffuse scattering sources has been subtracted from all
scans. Note that both as-quenched alloys yield essentially q-independent
behavior within the experimental statistical errors. This suggests that both
as-quenched alloys are homogeneous on a nanometer length scale. After the
430C anneal, both samples show a steep rise at the lowest q and this can be
fitted by a term of the form A / q3, where A is a constant, (solid line in
Fig. 5.44) corresponding to Porod behavior for a line-focus SAXS system
(Glatter and Kratky, 1982). That is, some large-scale electron density
fluctuations have developed even though no crystallization has occurred based
on the DSC, XRD, and Mossbauer results presented earlier. Such fluctuations
may be related to composition fluctuations detected by small-angle neutron
254 Michat Kopcewicz

scattering in similar Fe-based glasses and discussed in terms of fractal


concepts (Lamparter and Steeb, 1988) Detailed study requires
0

measurements to much lower q than accessible here 0

1500
FeSI Zr 7B 12

i0 1000
0

.
As-quenched
430C
0;;;
<= 0 600'C
~
oS
VJ

~
VJ
500

o 2 3

1500

1000 o As-quenched
~
~ .. 430C
0 o 600C
0;;;
<=
~
oS 500
VJ

~
VJ

o
o

o 2 3

Figure 5.44 SAXS data for Fe81 Zr7 8 12 and Fe79 Zr7 8 12 CU2alloys in the as-quenched state
and after annealing at 430 and 600C.

There is a clear and dramatic difference in the SAXS from the two types of
alloys annealed at 600 'c as displayed in Fig 0 50440 Although both alloys show
stronger scattering than after the 430 'c anneal, the Cu-containing alloy shows
significantly larger scattering intensities in the range up to q = 2 nm- 1 Fits to 0

the two sets of data were sought with a simple superposition of intensities from
spherical scattering objects using the known single-particle intensity function
for spheres (Glatter and Kratky, 1982) and including a Porod term (A / q3) to
account for the rise in intensity at low q due to larger, unresolved scattering
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 255

features, similar to those seen after the 430 'c anneal. The diameters of
spheres needed to fit the FeZrB were predominantly near 6 nm to 10 nm, but
there was some evidence for sizes below 2 nm. The smaller sizes have much
less weight in the SAXS intensities due to a proportionality of the intensity to
the square of the diameter (tor line-focus) .
The shape of the SAXS curve from the FeZrBCu alloy could not be fitted
with a superposition of spherical particle intensity functions which are known to
approach simple Gaussian shape at low q (Glatter and Kratky, 1982). The
clear "shoulder" in the data is evidence for an interparticle interference effect
in the scattering. This typically occurs for high densities of scattering centers
and a rather narrow size distribution (Glatter and Kratky, 1982). Such
interference effects can be modelled by including a pair correlation function
S ( q) such as that from the hard-sphere Percus-Yevick solution, shown to be
useful for precipitating alloys (Triolo et al., 1989). The solid line through the
data in Fig. 5.45 was obtained from this model for a hard-sphere pair
correlation function corresponding to particles of 2.8 nm average diameter but
with a hard-sphere" interaction" diameter of 6 nm with a volume fraction of
10%. This leads to an average distance between the centers of spheres of
about 13 nm. The difference between the particle and hard-sphere diameter is
discussed below.
2000
o As-quenched
o #3500'C
#3550'C
c #4500'C
1500
#4550'C

i<7
Vi 1000
"
2:l
.s
if)
X

if)

500

o 2 3
q (nm- I )

Figure 5.45 SAXS data for the as-quenched and annealed at 500 and 550'C samples #
3 Cx=8, y=O) and #4 Cx=8, y=2).

To interpret the origin of the SAXS, we assume the scattering to be due to


256 Michat Kopcewicz

the nanocrystalline bcc-Fe particles seen by the XRD and M6ssbauer


measurements after the 600 C anneal. For a two-phase mixture with a
difference in electron density Don between the two phases, the integrated
SAXS intensity can be used to determine the volume fraction , of the second
phase (Glatter and Kratky, 1982; Williamson, 1995). For line-focus
conditions and the intensity / ( q) in electron units,
'(1 - f) = Q/[2n2<O)(Don)2] (5.17)

Q = (Qsl2)fQ/(Q)dq (5.18)

where 'Qs is a geometrical factor determined by the slit system, <0) is the
average atomic volume of the sample, and the integration limits are, in
principle, from 0 to 00. In applying Eq. (5. 17) we include only the fitted
scattering contribution from the spheres (i. e., the Porod term is removed).
Values of Q from the FeZrB and FeZrBCu samples after the 600C anneal are
3.2 x 10 23 and 13. 8 x 1023 eu/cm 3 , respectively. In order to estimate volume
fractions we first calculated an electron density for the amorphous phase based
on atomic volumes of pure fcc Fe, hcp Zr, rhombohedral B, and fcc Cu, but
mixed with the nominal compositions Fes1 Zr7 B 12 and Fe79 Zr7 B 12 CU2' This yields
2.01 x 10 24 and 2.02 X 1024 cm- 3 , respectively, compared to 2.21 X 10 24 cm- 3
for pure bcc Fe. With the above Q' sand Eq. (5. 17) we first estimate' 0.03 =
=
and' 0.21 as the volume fraction of bcc-Fe in the form of nanoparticles for
the Cu-free and Cu-containing alloy, respectively. Since some of the Fe must
be segregated from the amorphous phase to produce the bcc-Fe particles, the
electron density of the amorphous phase will be lower than the above values
based on this model. We then search for the amount of Fe depletion needed to
reduce the electron density of the amorphous matrix and self-consistently yield
this same amount of Fe as the volume fraction f of pure bcc-Fe from
Eq. (5.17). This leads to densities of 1.97 x 1024 and 1.89 x 1024 cm- 3 and to
final estimates of ,= O. 02 and ,= O. 06 for the Cu-free and Cu-containing
alloys, respectively. The latter value is consistent with the above hard-sphere
interference model which yielded a 10% volume fraction of hard spheres. We
expect the hard-sphere volume fraction to be larger than the particle volume
fraction since the hard-sphere diameter represents the size of the Fe depletion
zone around each bcc-Fe particle. Indeed, the average bcc-Fe particle size
from fitting the data shown in Fig. 5. 44 for this sample is 2. 8 nm, compared to
the hard-sphere interaction diameter of 6 nm, also consistent with the
depletion zone concept.
The XRD, M6ssbauer, and SAXS data are all qualitatively consistent in
that smaller bcc-Fe grains are present and a larger fraction of the bcc-Fe is
produced for a given anneal temperature of the FeZrBCu alloy versus the
FeZrB alloy. However, there appear to be some quantitative inconsistencies
in the data regarding grain sizes and volume fractions. We will focus on the
data obtained after the 600C anneal since data from all three techniques are
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 257

available.
Analysis of bcc-Fe (110) XRD linewidths via the Scherrer formula yielded
grain sizes of 21 and 9 nm for the FeZrB and FeZrBCu, respectively, (see
Table 5.3) while analysis of the SAXS data provided average diameters about
a factor of two smaller for both alloys (8 and 2. 8 nm, respectively). The
SAXS provides information for the entire volume of the sample since it is a
transmission method while the XRD probes only 1 or 2 IJm into the surface
(based on a calculated Iinear absorption coefficient of 1530 cm - 1 for these
alloys) at the Bragg angle for the Fe (110) peak. Thus, more extensive grain
growth in the near-surface regions of the foils could account for this difference.
As it was shown earlier, the crystallization of amorphous ribbons starts indeed
at the surface (see Section 5.6). Another possibility is a wide or bimodal
distribution of bcc-Fe grain sizes such that the larger sizes are beyond the
resolution capability of the SAXS system and are therefore not detected except
as the Porod tail (i.e., the q-3 behavior at the lowest q's) and therefore not
included in the average size analysis via SAXS. A bimodal type distribution
would be more consistent with the SAXS data from the FeZrBCu alloy since the
observed interference effect (shoulder in SAXS curve) is usually indicative of a
rather narrow size distribution. Then, the larger size part of the bimodal
distribution would have to be >20 nm to avoid detection in the SAXS except as
a q-3 term.
The volume fractions of the bcc-Fe nanoparticles detected by SAXS
(about 2 % and 6 % for the FeZrB and FeZrBCu alloys after 600 'c,
respectively) are much smaller than those indicated by the fractional Fe
Mossbauer resonance (Fig. 5. 13a, about 10% and 45%, respectively). The
latter values need to be converted into volume fractions and corrected for
possible differences in recoilless fractions between the amorphous and bcc-Fe
phases to enable a direct comparison. However, this will not account for the
discrepancy and a possible error in the calculated electron densities for the
amorphous phase used in Eq. (5.17). That is, a smaller density contrast l::.n
may be correct so that larger volume fractions are present. Also, as above, a
bimodal grain size distribution may be responsible such that a significant
volume fraction of bcc-Fe is not being detected by SAXS due to much larger
sizes. The rf-Mossbauer data (Figs. 5. 20 and 5. 21) provided indirect
evidence of a bimodal grain distribution in the FeZrB alloy due to a mixture of
magnetically hard and soft bcc Fe; however, similar evidence was not found
for the FeZrBCu alloy where the data suggested rather a more continuous size
distribution.
The SAXS measurements were performed for all Fe93-x-yZr7BxCuy alloys
annealed at 500 and 550 'c (Kopcewicz et al., 1997a). Examples for
Fea5 Zr7 Baand Fea3 Zr7 BaCU2 are shown in Fig. 5.45 compared to the data from
an as-quenched sample. The experimental procedure and method of fitting the
SAXS data are the same as before. All scans show a constant background
intensity due to the Fe fluorescence and other diffuse scattering. All scans,
258 Michal Kopcewicz

except those from as-quenched samples, show a steep rise of the SAXS
intensity for very low momentum transfers (q), which can be fitted by a term
of the form A / q3 (A is a constant) corresponding to Porod behavior for a line-
focus SAXS system as used here. This Porod tail is related to large-scale
(~20 nm) electron density fluctuations.
The SAXS patterns differ considerably for the two groups of alloys: that
with no Cu, and the other containing 2 % Cu. All Cu-containing alloys show
=
significantly larger scattering intensities in the range up to q 2 nm - 1 , at the
same annealing temperatures, which increase with increasing T A (Fig. 5.45).
The fits to the data were performed with the superpositions of intensities from
spherical scattering objects and a Porod term (A / q3) to account for the rise of
the intensity at low q together with a constant background intensity. The
average diameters of spheres obtained from the fits are included in Table 5.3
for convenient comparison to the XRD results. The shape of SAXS data
obtained for the Cu-containing alloys annealed at 550 'c is more complex. A
clear "shoulder" in the data (Fig. 5.45, sample :1:1: 3, see Table 5.3) is
evidence for an interparticle interference effect in the scattering which occurs
for a high density of scattering centers with a fairly narrow size distribution
(Glatter and Kratky, 1982). Such an interference effect can be modelled by
including a pair of correlation function S ( q) based on the hard-sphere Percus-
Yevick solution for precipitating alloys (Triolo et aI., 1989).
To interpret the origin of the SAXS, we assume that the scattering is due
to the nanocrystalline bee-Fe particles seen by XRD and Mossbauer
measurements. For a two-phase mixture the integrated SAXS intensity Q is
proportional to the volume fraction of nanoparticles. Values of Q are listed in
Table 5.3. In order to calculate the volume fraction it is necessary to calculate
an electron density for the amorphous phase as suggested above. Here, rather
than attempt to estimate absolute values of volume fractions, the SAXS
intensities, Q' s, in Table 5. 3 can be used to follow the relative volume
fractions versus alloy composition (x and y).
The SAXS results provided detailed information regarding size of the bcc-
Fe grains. The grains are smaller in the Cu-containing alloys with same boron
content at a given anneal ing temperature. This result is in good qual itative
agreement with the XRD data. The number of nanocrystals (volume fraction of
the bee-Fe grains) increases with anneal temperature for each set of x and y
(Q in Table 5. 3). If we regard the coefficient of the Porod term, A,
(Table 5.3), as a measure of larger-scale features, it does not vary
systematically with x and y or with anneal temperature. This term is likely due
in part to the larger grains near the surface as detected by XRD and in part due
to composition fluctuations that have nothing to do with crystallization. The
average grain sizes <D) determined by SAXS are approximately a factor of
two smaller than those resulting from XRD L ) ) .
This difference between SAXS and XRD data may result from the fact that
in SAXS, which is a transmission technique, the entire volume (""" 20 IJm
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 259

thick) of the sample is probed, while in XRD, which is a scattering technique.


only a surface layer is probed whose thickness is determined by the absorption
coefficient of the alloy (~2200 cm- 1 ) and the Bragg angle of a given Fe peak.
For example, 50 % of the X-ray signal comes from a layer only O. 6 IJm thick
for the (110) diffraction peak. The different development of the grains at the
surface and in the bulk could therefore account for the difference in average
sizes found by the two methods. Clear differences between surface and bulk
crystallization of FeZrB (Cu) alloys were detected recently by conversion
electron M6ssbauer spectroscopy. It was found that the crystallization is more
pronounced at the surface as compared to the bulk at a given annealing
temperature, and in some cases different phases are formed (see Section
5.6, Figs. 5.31-5.34).
The SAXS results show good qualitative agreement with the M6ssbauer
data. The volume fraction of bee grains determined by SAXS (parameter Q,
Table 5.3) increases with annealing temperature for each set of x and y in
good agreement with M6ssbauer results shown in Figs. 5. 27 and 5. 28. The
finding that the bee grains, as determined by SAXS (Table 5.3), are smaller
in Cu-containing alloys at a given boron content and the same annealing
temperature corresponds well to the conclusion drawn from the rf-M6ssbauer
investigations, which reveal that the magnetic anisotropy and, hence the grain
sizes are smaller in the Cu-containing alloys. However. the bimodal shape of
the grain size distribution inferred from the rf-M6ssbauer spectra of the FeZrB
alloys can not be confirmed by SAXS, probably because the accessible
q-range at low q is not sufficient. The steep rise in the data at low q is
evidence of some larger-scale structure, the origin of which is not yet clear. It
is interesting to note that the SANS experiment on the crystallization of the
Fe80 B20 alloy (Deriu et aI., 1996) found evidence of a bimodal size distribution
of bee-Fe nanocrystals by accessing lower q' s than used here. The larger
sizes detected were near 90 nm.
The SAXS and M6ssbauer techniques complement each other well. The
SAXS technique allows a direct determination of the absolute value of the
average size of nanoparticles without providing information regarding magnetic
properties. whereas the rf-M6ssbauer technique delivers information on the
magnetic properties (anisotropy, magnetostriction) from which one can infer
qualitative conclusions regarding the grain size distribution. Perhaps
theoretical work leading to precise fitting of the rf-M6ssbauer spectra will
eventually allow absolute grain size information to be extracted. Finally, we
note that both SAXS and the M6ssbauer techniques as appl ied here are
transmission methods, so they probe the same entire volume of the sample. in
contrast to the XRD method, which probes a relatively small volume of the
sample near the surface.
260 Michat Kopcewicz

References
Albanese, G., G. Asti and S. Rinaldi. Nuevo Cimento B 6: 153(1971)
Albanese, G., G. Asti and S. Rinaldi. Nuevo Cimento Lett. 4: 220( 1972)
Alben, R., J. J. Becker and M. C. Chi. J. Appl. Phys. 49: 1653(1978)
AIIia, P., M. Baricco, P. Tiberto and F. Vinai. J. Appl. Phys. 74: 3137
(1993)
Asti, G., G. Albanese and C. Bucci. Phys. Rev. 184: 260 ( 1969)
Barandiaran, J. M., P. Gorria, I. Orue, M. L. Fernando-Gubieda, F.
Plazaola and A. Hernando. Phys. Rev. B 54: 3026 ( 1996)
Borrego, J. M., C. F. Conde, A. Conde, V. A. Pena-Rodriguez and J. M.
Greneche. J. Phys. C 12: 8089(2000)
Clavaguera, N. , T. Pradell, Z. Jie and M. T. Clavaguera-Mora.
Nanostructured Mater. 6: 453 ( 1995)
Cserei, A., J. Jiang, F. Aubertin and U. Gonser. J. Mater. Sci. 29: 1213
(1994)
Deriu, A., F. Malizia, F. Ronconi, M. Vittori-Antisari and S. M. King. J.
Appl. Phys. 79: 2296( 1996)
Duhaj, P., I. Matko, P. Svec, J. Sitek and D. Jani6kovi6 Mater. Sci. Eng.
B 39: 208(1996)
Girchardt, T., T. Graf, oj. Hesse, A. Grabias, M. Kopcewicz and G.
Herzer. Mater. Sci. Eng. A 226 - 228: 204 (1997)
Girchardt, T., J. Hesse, A. Grabias, M. Kopcewicz, D. Ramin and W.
Riehemann. Nanostructured Mater. 12: 93l( 1999)
Glatter, O. and O. Kratky. Small Angle X-ray Scattering. Academic, London
(1982)
Gomez-Polo, C., D. Holzer, M. Multigner, E. Navarro, P. Agudo, A.
Hernando, M. Vazquez, H. Sassik and R. Grossinger. Phys. Rev. B 53:
3392(1996)
Gonser, U., M. Ackermann and H. G. Wagner. J. Magn. Magn. Mater. 31-
34: 1605(1983)
Gonser, U. and H.G. Wagner. Hyperfine Inter. 24-26: 769(1985)
Gorria, P., I. Orue, F. Plazaola, M. L. Fernandez-Gubieda and J. M.
Barandiaran. IEEE Trans. Magn. 29: 2682(1993)
Gorria, P., J. S. Garitaonandia and J. M. Barandiaran. J. Phys. C 8: 5925
(1996)
Grabias, A. and M. Kopcewicz. Mater. Sci. Forum 269-272: 725(1998)
Grabias, A., M. Kopcewicz and B. Idzikowski. Mater. Res. Soc. Symp.
Proc. 577: 543(1999a)
Grabias, A., M. Kopcewicz and B. Idzikowski. Nanostructured Mater. 12:
899(1999b)
Grabias, A., M. Kopcewicz and B. Idzikowski. Hyperfine Inter. 126: 21
(2000)
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 261

Grat. T., M. Kopcewicz and J. Hesse. J. Magn. Magn. Mat. 140-144: 423
(1995a)
Grat, T., M. Kopcewicz and J. Hesse. Nanostructured Mater. 6: 937(1995b)
Grat, T., M. Kopcewicz, J. Hesse. J. phys.: Condens. Matter 8: 3897
(1996)
Grat, T., J. Hesse and M. Kopcewicz. Nanostructured Mater. 12: 935 ( 1999)
Greneche, J. M. and M. Miglierini. In: M. Miglierini and D. Petridis eds.
Mossbauer Spectroscopy in Materials Science, Kluwer, Dordrecht, p. 243
(1999)
Gupta, A., N. Bhagat and G. Principi. J. Phys. C 7: 2237( 1995)
Hampel, G., A. Pundt and J. Hesse. J. Phys. C 4: 3195( 1992)
Hanna, S. S., J. Heberle, C. Littlejohn, G. J. Perlow, R. S. Preston and D.
H. Vincent. Phys. Rev. Lett. 4: 28 (1960a)
Hanna, S. S., J. Heberle, C. Littlejohn, G. J. Perlow, R. S. Preston and D.
H. Vincent. Phys. Rev. Lett. 4: 177(1960b)
Hendricks, R.W., P.G. MardonandL.B. Schaffer. J. Chem. Phys. 61: 319
(1974)
Hernando, A. and T. Kulik. Phys. Rev. B 49: 7(1994)
Hernando, A., M. Vazquez, T. Kulik and C. Prados. Phys. Rev. B 51: 3581
(1995)
Herzer, G. IEEE Trans. Magn. 26: 1397 ( 1990)
Herzer, G. Mater. Sci. Eng. A 133: 1(1991)
Herzer, G. Physica Scripta T49: 307(1993)
Hesse, J. and A. Ri.ibartsch. J. Phys. E 7: 526( 1974)
Hoving, W., F. Van der Woude and K. H. J. Buschow. In: S. Steeb and H.
Warlimont eds. Proc. Fifth Int. Conf. on Rapidly Quenched Metals. North-
Holland, Amsterdam, p. 549 (1985)
Jiang, J., T. Zemcik, F. Aubertin, u. Gonser. J. Mat. Sci. Lett. 10: 763
( 1991)
Kamiyama, T., S. Matsuo and K. Suzuki. J. Non-Cryst. Solids 150: 172
(1992)
Kataoka, N., M. Hosokawa, A. Inoue and T. Masumoto. Japan J. Appl.
Phys. 28: L262 (1989)
Kemeny, T., J. Balogh, I. Farkas, D. Kaptas, L. F. Kiss, T. Pusztai, L.
Toth and I. Vincze. J. Phys. C 10: L221(1998)
Kemeny, T., D. Kaptas, J. Balogh, L.F. Kiss, T. Pusztai and I. Vincze. J.
Phys. ell: 2841 ( 1999)
Keune, W., R. Halbauer, U. Gonser, J. Lauer and D. L. Williamson. J.
Appl. Phys. 48: 2976( 1977)
Kim, K. Y., T. H. Noh, I. K. Kang and T. Kang. Mater. Sci. Eng. A 179/
180: 552(1994)
Kim, C.S., S. B. Kim, J. S. Lee and T. H. Noh. J. Appl. Phys. 79: 5459
(1996)
Kistner, O. C. and A. W. Sunyar. Phys. Rev. Lett. 4: 412( 1960)
262 Michat Kopcewicz

Klug, H. P. and L. E. Alexander. X-ray Diffraction Procedures, 2nd. Wiley,


New York, p. 618(1974)
Kobayashi, H., H. Onodera and H. Yamamoto. J. Phys. Soc. Japan 55: 331
(1986)
Kohlbrecher, J., A. Wiedenmann and H. Wollenberger. Physica B 213/214:
579( 1995)
Kopcewicz, M., A. Kotlicki and M. Szefer. Phys. Status Solidi (b) 72: 701
(1975 )
Kopcewicz, M. Phys. Status Solidi (a) 46: 675(1978)
Kopcewicz, M. and A. Kotlicki. J. Phys. Chem. Solids 41: 631 (1980)
Kopcewicz, M., H. G. Wagner and U. Gonser. J. Magn. Magn. Mater. 51:
225(1985)
Kopcewicz, M., H.G. Wagner and U. Gonser. Hyperfine Inter. 27: 413
( 1986a)
Kopcewicz, M., H.G. Wagner and U. Gonser. J. Phys. F 16: 929(1986b)
Kopcew icz , M. In: G. J. Long and F. Grandjean eds. Mossbauer
Spectroscopy Applied to Inorganic Chemistry. Plenum Press, New York-
London, Vol. 3, p. 243 (1989)
Kopcewicz, M. Structural Chem. 2: 313 (1991)
Kopcewicz, M., B. Idzikowski and A. Wrzeciono. J. Magn. Magn. Mater.
125: 290(1993)
Kopcewicz, J. Jagielski, T. Graf, M. Fricke and J. Hesse. Hyperfine Inter.
94: 2223 (1994a)
Kopcewicz, M., J. Jagielski, T. Stobiecki, F. Stobiecki and G. Gawlik. J.
Appl. Phys. 76: 5232( 1994b)
Kopcewicz, M., A. Grabias and P. Nowicki. J. Magn. Magn. Mat. 140-
144: 461 (1995a)
Kopcewicz, M., A. Grabias and P. Nowicki. Nanosctructured Mater. 6: 957
(1995b)
Kopcewicz, M., A. Grabias and P. Nowicki. Mater. Res. Soc. Symp. Proc.
384: 517( 1995c)
Kopcewicz, M., B. Idzikowski, J. Kovac and A. Wrzeciono. J. Magn.
Magn. Mater. 140-144: 315(1995d)
Kopcewicz, M., A. Grabias, P. Nowicki and D. L. Williamson. J. Appl.
Phys. 79: 993(1996)
Kopcewicz, M. and A. Grabias. J. Appl. Phys. 80: 3422( 1996)
Kopcewicz, M., A. Grabias and D. L. Williamson. J. Appl. Phys. 82: 1747
(1997a)
Kopcewicz, M., A. Grabias and P. Nowicki. Mater. Sci. Eng. A 226-228:
515 (1997b)
Kopcewicz, M., A. Grabias and P. Nowicki. In: P. Duhaj, M. Mrafko and P.
Svec eds. Proc . Int. Cant. on Rapidly Quenched and Metastable
Materials, Elsevier, Amsterdam, p. 549 (1997c)
Kopcewicz, M., A. Grabias and B. Kopcewicz. J. Magn. Magn. Mater.
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 263

177-181: 73(1998a)
Kopcewicz, M., A. Grabias and I. Skorvanek. J. Appl. Phys. 83: 935
( 1998b)
Kopcewicz, M., A. Grabias and B. Idzikowski. Mater. Res. Soc. Symp.
Proc. 577: 487 ( 1999a)
Kopcewicz, M., A. Grabias and B. Idzikowski. Mater. Res. Soc. Symp.
Proc. 577: 563(1999b)
Kopcewicz, M., A. Grabias, I. Skorvanek, I. Marcin and B. Idzikowski. J.
Appl. Phys. 85: 4427 (1999c)
Kopcewicz, M., A. Grabias, B. Idzikowski and D. L. Williamson. Hyperfine
Inter. 139/140: 525(2002)
Kopcewicz, M., A. Grabias, M. A. Willard, D. E. Laughlin and M. E.
McHenry. IEEE Trans. Magn. 37: 2226(2001)
Kulik, T., A. Hernando and M. Vazquez. J. Magn. Magn. Mater. 133: 310
(1994)
Lamb, W.E. Jr. Phys. Rev. 55: 190(1939)
Lamparter, P. and S. Steeb, J. Non-Cryst. Solids 106: 137(1988)
LeCaer, G. and J.M. Dubois. J. Phys. E 12: 1083(1979)
Makino, A., A. Inoue and T. Masumoto. Nanostructured Mater. 6: 985 ( 1995)
Makino, A., T. Hatanai, A. Inoue and T. Masumoto. Mater. Sci. Eng.
A 226 - 228: 594( 1997a)
Makino, A., T. Hatanai, Y. Naitoh, T. Bitoh, A. Inoue and T. Masumoto.
IEEE Trans. Magn. 33: 3793( 1997b)
Makino, A., A. Inoue and T. Masumoto. Mater. Res. Soc. Symp. Proc.
577: 457( 1999)
McHenry, M. E., M. A. Willard and Laughlin. Progr. Mater. Sci. 44: 291
(1999)
Miglierini, M. J. Phys. C 6: 1431(1994)
Miglierini, M., J. Lipka and J. Sitek. Hyperfine Inter. 94: 2193(1994a)
Miglierini, M., J. Sitek, Z. Szasz and K. Vitazek. Hyperfine Inter. 84: 295
(1994b)
Miglierini, M. and J. M. Greneche. J. Phys. C 9: 2303(1997a)
Miglierini, M. and J. M. Greneche. J. Phys. C 9: 2321(1997b)
Miglierini, M. Y. Labaye, N. Randrianantoandro and J. M. Greneche.
Mater. Sci. Eng. A 226 - 228: 559 ( 1997)
Miglierini, M., I. Skorvanek and J. M. Greneche. J. Phys. C 10: 3159
(1998)
Miglierini, M. and J. M. Greneche. Hyperfine Inter. 113: 375(1998)
Miglierini, M., M. Kopcewicz, B. Idzikowski, Z. E. Horvath, A. Grabias, I.
Skorvanek, P. Dluzewski and Cs. S. Daroczi. J. Appl. Phys. 85: 1014
( 1999)
Miglierini, M. and J. M. Greneche. In: M. Miglierini and D. Petridis eds.
Mossbauer Spectroscopy in Materials Science. Kluwer, Dordrecht,
p. 257 (1999a)
264 Michat Kopcewicz

Miglierini, M. and J. M. Greneche. Hyperfine Inter. 122: 121(1999b)


Miglierini, M. and J. M. Greneche. Hyperfine Inter. 121/122: 297( 1999c)
Miglierini, M., J. M. Greneche and B. Idzikowski. Mater. Sci. Eng. A 304-
306: 937 (2001)
Mbssbauer, R. L. Z. Phys. 151: 124( 1958a)
Mbssbauer, R. L. Naturwissenschaften 45: 538 (1958b)
Muller, M, N. Mattern and L. IIlgen. J. Magn. Magn. Mater. 112: 263
(1992)
Navarro, I., A. Hernando, M. Vazquez and S. C. Yu. J. Magn. Magn.
Mater. 145: 313(1995)
Ohnuma, M., J. Suzuki, S. Funahashi, T. Ishigaki, H. Kuwano and Y.
Hamaguchi. Physica B 213/214: 582 ( 1995)
Ok, H.N. and A.H. Morrish. Phys. Rev. B 23: 2257(1981)
Orue, I., P. Gorria, F. Plazaola, M. L. Fernandez-Gubieda and J. M.
Barandiaran. Hyperfine Inter. 94: 2199 ( 1994)
Pfeiffer, L. In: I. J. Gruvermann ed. Mossbauer Effect Methodology, Vol. 7.
Plenum Press, New York, Vol. 7, p. 263 (1972)
Pfeiffer, L., N.D. Heiman and J.C. Walker. Phys. Rev. B 6: 74(1972)
Pradell, T., N. Clavaguera, J. Zhu and M. T. Clavaguera-Mora. J. Phys. C
7: 4129( 1995)
Pundt, A., G. Hampel and J. Hesse. Z. Phys. B 87: 65( 1992)
Randrianantoandro, N., J. M. Greneche, E. Jedryka, A. Slawska-Waniewska
and H. K. Lachowicz. Mater. Sci. Forum 179 - 181: 545 (1995)
Randrianantoandro, N., A. Slawska-Waniewska and J. M. Greneche. Phys.
Rev. B56: 74(1997)
Rixecker, G., P. Schaaf and U. Gonser. J. Phys. C 4: 10295(1992)
Saegusa, N. and A.H. Morrish. Phys. Rev. B 26: 6547(1982)
Schiffer, J. and W. Marshall. Phys. Rev. Lett. 3: 556 (1959)
Skorvanek, I. and M. Miglierini. J. Magn. Magn. Mater. 96: 162(1991)
Skorvanek, I. and R. Gerling. J. Appl. Phys. 72: 3417(1992)
Skorvanek, I., M. Miglierini and P. Duhaj. Mater. Sci. Forum 235 - 238: 771
(1997a)
Skorvanek, I., P. Duhaj, J. Kovac, V. Kavecansky and R. Gerling. Mater.
Sci. Eng. A 226-228: 218(1997b)
Skorvanek, I., J. Kovac and J. M. Greneche. J. Phys. C 12: 9085(2000)
Slawska-Waniewska, A., M. Gutowski, H. K. Lachowicz, T. Kulik and H
Matyja. Phys. Rev. B 46: 14,594(1992)
Slawska-Waniewska, A., A. Roig, E. Molins, J. M. Greneche and R.
Zuberek. J. Appl. Phys. 81: 4652(1997)
Slawska-Waniewska, A. and J. M. Greneche. Phys. Rev. B 56: R8491C 1997)
Srivastava, J. K. In: B. V. Thosar, P. K. Iyengar, J. K. Srivastava and S.
C. Bhargava eds. Advances in Mossbauer Spectroscopy. Elsevier,
Amsterdam, p. 761 (1983)
Mossbauer Spectroscopy Characterization of Soft Magnetic. . . 265

Sui, M. L., K. Y. He, L. Y. Xiong, Y. Liu and J. Zhu. Mater. Sci. Eng. A
181/A182: 1405(1994)
Suzuki, K., N. Kataoka, A. Inoue, A. Makino and T. Masumoto. Mater.
Trans. JIM 31: 743( 1990)
Suzuki, K., A. Makino, N. Kataoka, A. Inoue and T. Masumoto. Mater.
Trans. JIM 32: 93(1991a)
Suzuki, K., A. Makino, A. Inoue and T. Masumoto. J. Appl. Phys. 70: 6232
(1991b)
Suzuki, K., A. Makino, A. InoueandT. Masumoto. J. Appl. Phys. 74: 3316
(1993)
Suzuki, K., A. Makino, A. P. Tsai, A. Inoue and T. Masumoto. Mater. Sci.
Eng. A 179 - 180: 501( 1994)
Suzuki, K. and J.M. Cadogan. Phys. Rev. B 58: 2730(1998)
Suzuki, K. and J. M. Cadogan. J. Appl. Phys. 85: 4400 ( 1999)
Takacs, L., M.C. Cadeville and I. Vincze. J. Phys. F 5: 800(1975)
Triolo, R., E. Caponetti and S. Spooner. Phys. Rev. B 39: 4588 ( 1989)
Wagner, H. G., M. Ackermann, R. Gaa and U. Gonser. In: S. Steeb and H.
Warlimont eds. Proc. Fifth Int. Conf. on Rapidly Quenched Metals, North-
Holland, Amsterdam, p. 247 (1985)
Willard, M. A., D. E. Laughlin, M. E. McHenry, D. Thoma, K. Sickafus, J.
O. Cross and V. G. Harris. J. Appl. Phys. 84: 6773 ( 1998)
Willard, M. A., M. -Q. Huang, D. E. Laughlin, M. E. McHenry, J. O. Cross,
V.G. Harris and C. Franchetti. J. Appl. Phys. 85: 4421(1999)
Williamson, D. L., A. H. Mahan, B. P. Nelson and R. S. Crandall. Appl.
Phys. Lett. 55: 783( 1989)
Williamson, D.L. Mater. Res. Soc. Symp. Proc. 377: 251(1995)
Yang, X. L., J. X. Yang, K. Y. Jiang, G. Chen, H. J. Jin and Y. Z. Zhang. J.
Appl. Phys. 85: 5124(1999)
Yoshizawa, Y., S. Oguma and K. Yamauchi. J. Appl. Phys. 64: 6044 ( 1988)
Yoshizawa, Y., Y. Bizen and S. Arakawa. Mater. Sci. Eng. A 181/A182:
871(1994)
Zemcik, T., K. Zaveta and I. SkorvC:mek. J. Mater. Sci. 28: 654 ( 1993)
Zhou, X. Z., A. H. Morrish, D. G. Naugle and R. Pan. J. Appl. Phys. 73:
6597( 1993)

The financial support from the Grant No. KBN 2 P03B 068 15 and 3 P03B 07524 from the
State Committee of Scientific Research (Poland) is gratefully acknowledged.
6 Atom Probe Characterization of Microstructures
of Nanocrystalline and Nanocomposite
Magnetic Materials

Kazuhiro Hono

6. 1 Introduction

Both nanocrystalline soft and hard magnetic materials are relatively new series
of magnetic alloys. The former was invented by Yoshizawa et al. in 1988 by
crystallizing the Fe-Si-B based amorphous alloy microalloyed with Cu and Nb
(Yoshizawa et aI., 1988). Due to exchange-coupled randomly oriented
nanograins, the magnetocrystalline anisotropy of the Fe-based phase is
averaged out, resulting in a low coercivity. The nanocomposite hard magnetic
material was also reported in 1988 by Coehoon et al. (1988). The
microstructure was composed of nanoscale soft Fe3 B grains and hard Nd2 Fe14 B
grains, which are produced from Fe-Nd-B based alloy either by crystallizing
during rapid solidification or from the overquenched amorphous precursors.
When the grain size is less than 50 nm, remanence is significantly enhanced
from the limit of the isotropic polycrystalline magnet (M, > Mj2, here M, is
remanent magnetization and M s is saturation magnetization) due to the
remanence enhancement effect from exchange-coupled soft and hard
nanograins. In both cases, the nanocrystalline microstructure is produced via
the crystallization route from the amorphous precursors that are produced by
rapidly solidifying melts; thus understanding the mechanisms of the
nanocrystalline microstructural evolution is essential to control the magnetic
properties of nanocrystalline soft and hard magnetic alloys.
Refinement of crystal grain size in less than 20 nm is necessary to obtain
low coercivity in nanocrystalline soft magnetic alloys as well as to obtain high
remanence and high energy product in nanocomposite magnets. On this basis,
many attempts were made to improve the magnetic properties of
nanocrystalline and nanocomposite magnets by adding quaternary and quinary
elements in base ternary alloy compositions. However, most of these
elements were added to the base alloy based on the prior experience obtained
by trial and error without much understanding of the underlying mechanisms. If
we understand how these additives work in the microstructural refinement as
well as the magnetic property improvement, it should be useful for designing
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 267

high performance nanocrystall ine magnetic materials more efficiently. For this
purpose, it is necessary to observe how microalloyed elements distribute in
the initial stage, and how redistribution of the elements proceeds during the
nanocrystalline microstructural evolution.
However, there are two difficulties in locating microalloyed elements in
the nanocrystalline and nanocomposite microstructures, i. e., CD the size of the
microstructure is so small that most of the analytical techniques do not have
sufficient spatial resolution to analyze the chemical compositions of the
nanoscale particles embedded in a matrix phase; ~ although the partitioning
and redistribution behaviors of light elements play critical roles in the
nanocrystalline microstructure evolution, most of the analytical techniques are
not capable of detecting light elements in a nanoscale dimension
quantitatively. Unlike other techniques, the three-dimensional atom probe
(3DAP) does not receive such restrictions. It detects individual atoms from the
surface of a specimen, determ ining both mass and positions of individual
atoms. Thus, it is possible to determine local compositions of the nanoscale
particles embedded in a matrix phase by counting the atoms exclusively from
the nanoparticles. It can also detect light elements with equal detection
efficiency; hence, quantitative analysis of light elements from subnanoscale
volume can be done. These are the reasons why the atom probe (AP)
technique has been successful in characterizing the microstructures of
nanocrystalline and nanocomposite microstructures in the past ten years.
This chapter gives a brief introduction of the modern AP technique, and
then gives an overview on its applications to the studies on the nanocrystalline
and nanocomposite microstructural evolution processes in Fe based soft and
hard magnetic materials. In addition, the roles of various microalloyed
elements in the nanocrystalline microstructure evolution will be discussed
based on recent atom probe studies. Key factors to control the nanocrystalline
and nanocomposite microstructures will also be summarized.

6.2 The Atom Probe Technique

The atom probe was originally invented by E. W. Muller and his coworkers in
1968 (Muller, 1968). It is a combination of a field ion microscope (FIM) and a
time-of-flight mass spectrometer. Using an atom probe, it is possble to detect
individual atoms that are ionized from sharp needle-like metallic specimens.
Atoms are ionized under a very high electric field applied on sharp metal tips
by the field evaporation process. This ionization occurs from the surface of the
specimen regularly, so it is possible to ionize atoms with an atomic layer
order, by which atomic layer resolution can be achieved on the atom probe
analysis. Two types of atom probes are currently used: one is a conventional
268 Kazuhiro Hono

time-of-flight atom probe. the schematic illustration of which is shown in


Fig. 6. la. By selecting a small region using an aperture called the probe
hole, it is possible to collect atoms only from the region selected by the probe
hole. The physical dimension of the probe hole is typically 2 to 3 mm in
diameter, but its projected size is less than 5 nm. Atoms are evaporated from
the surface continuously, thus it is possible to collect atoms in the three
dimensional volume covered by the probe hole. However, in this atom probe,
the position information of the atoms is not recorded Hence, the data chain of
the collected atoms can be converted to a one-dimensional depth profile by
scaling the depth in proportion to the number of collected atoms. Because of
the nature of the information obtained from this type of atom probe, the
conventional time-of-flight atom probe is now called the one-dimensional atom
probe (1 DAP) to differentiate it from the recently developed three-dimensional
atom probe (3DAP). The lateral resolution is determined by the size of the
probe hole and the evaporation aberration.

Probe hole
Detector
------I m.
---- [lq

Depth
(a) (b)

Figure 6. I Schematic illustrations for the principle of Ca) a conventional time-of-flight


atom probe CAP) and Cb) a three-dimensional atom probe C3DAP)

Unlike the 1DAP, the 3DAP obtains the information on the lateral position
of atoms using a position sensitive detector (PSD) as shown in Fig. 6. 1b.
This type of atom probe was originally developed as a position sensitive atom
probe (PoSAP) by Cerezo et al. (1988), followed by Blavette et al. (1993)
as a tomographic atom probe (TAP). Since the AP data detected by a PSD
give three-dimensional elemental maps, this type of AP is now generally called
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 269

a three-dimensional atom probe (3DAP). By measuring the time-of-flight and


the coordinates of ions using a PSD, it is possible to map out two-dimensional
elemental distribution with a near-atomic resolution. The lateral spatial
resolution is limited by the evaporation aberration that occurs during the
ionization of atoms on the surface. However, the error originated from the
evaporation aberration does not exceed 0.2 nm, and this is still the lowest of
the errors achieved by any existing analytical instruments. The elemental
maps can be extended to the depth direction by ionizing atoms from the surface
of the specimen continuously, by which atom distribution can be reconstructed
in a three-dimensional real space. Since field evaporation occurs layer-by-
layer in low index planes, the reconstructed 3D elemental map shows the
layers corresponding to the atomic planes in the depth direction. Data
visualization can be processed using a three-dimensional graphic software.
Although earlier 3DAP instruments could not achieve a high mass
resolution in the time-of-flight mass analysis, the modern 3DAP instruments are
equipped with a reflectron energy compensator, and are able to achieve a
mass resolution, m / Ilm, larger than 300 (Cerezo et al., 1998). With this
performance, most of the alloying elements contained in complicated
nanocrystalline magnetic alloys can be identified without any ambiguity. The
detailed account of the instrumentation as well as the data analysis techniques
are nicely presented in a recent text book on atom probe tomography (Miller,
2000) .
In order to demonstrate the unique analytical capability of the 3DAP
technique, typical data obtained from a nickel-base superalloy (Inconel X-750:
Ni-16. 8Cr-6. 5Fe-2. 9Ti-1. 5AI) is shown in Fig. 6. 2. Nickel-base superalloys
form nanoscale cuboidal precipitate with the L 12 structure, Nb (AI, Ti), and the
AI and Ti enriched regions in the elemental map correspond to these
precipitate particles. Figure 6. 2a shows a three-dimensional elemental map of
AI and Ti atoms in this alloy. From the number density of these solute atoms,
the morphology of Nb (AI, Ti) particles can be clearly seen. The chemical
compositions of these nanometer sized particles can be determined by counting
the number of solute atoms as a function of the total number of detected atoms
within the volume of the analysis as shown in Fig. 6. 2b. Since atoms are
counted exclusively from the particles, the composition calculated by this
method gives the true chemical compositions of the nanosized particles. In the
L 12 Ni 3 (AI, Tj) precipitates, AI and Ti occupy the AI sub lattice of the Nb AI
ordered phase. Corresponding to this structure, the atomic layers can be
resolved in the [001] direction within the particles as shown in Fig. 6. 2c. This
demonstrates that the 3DAP has an atomic resolution in the depth direction of
270 Kazuhiro Hono

- -,.} ~.,
-IOOnm
(a)

'Cr' Ni

0'20~Fe
".;::
0.15 .

I::~~.
u 0.15
0.10 Ti
Fe' Ni
0.05
o
O'IO~
0.08
~.~
006 _ _
AI
l"'kA_
fL
o 20 40 60 80 100
Depth (nm)
(b) (c)

Figure 6. 2 Typical 3DAP data obtained from an Inconel X-760 alloy


(Ni-16. 8Cr-6. 5Fe-2. 9Ti-l .5AI alloy: solution heat treated and aged for 75 h at 750'C .
(a) Mass-spectrum obtained using an energy compensated 3DAP; (b) AI and Ti atom map
within an analysis volume of approximately 15 x 15 x 40 nm; (c) Concentration depth
profile of each alloying element calculated from the selected volume of 7 x 7 x 40 nm;
(d) Enlarged elemental maps near a y / y' interface.

the analysis. Figure. 6. 2c is an enlarged elemental map of one of the y / y'


interfaces in Fig. 6. 2a, in which the elemental distribution near the interface is
atomically resolved. As demonstrated in this data, the 3DAP gives very
accurate compositional information on the interface of nanosized particles
embedded in a matrix phase. This type of analysis is impossible with an
analytical transmission electron microscope (AEM), because the typical thin
foil thickness of TEM (transmission electron microscope) specimens is larger
than 200 nm, in which several nanosized particles are embedded within the
matrix. So the energy dispersive X-ray spectra (EDXS) obtained from the
nanosized particles are always influenced by the matrix phase as .shown in
Fig. 6.3. This is the reason why the atom probe technique has definite merit
over AEM when it comes to the analysis of nanosized particles that are
precipitate out from the matrix phase.
Atom Probe Characterization of Microstructures of Nanocrystalline... 271

Electron beam
EDS

I \
I \
I \
I \
I \
I \

Figure 6. 3 Schematic illustration to show how energy dispersive X-ray spectrum is


obtained from a TEM specimen containing nanosized particles.

6.3 Microstructural Evolution in Nanocrystalline Soft


Magnetic Materials

6.3.1 FINEMET (Fe-Si-B-Nb-Cu)


In 1988, Yoshizawa et al. reported that excellent soft magnetic properties are
obtained when the crystal grain size is reduced to a nanometer scale by
crystallizing an Fe-Si-B-Nb-Cu amorphous alloy. Nanocrystallization itself has
been known for a long time, but this was the first report that nanocrystalline
ferromagnetic materials exhibit soft magnetic properties. A TEM image of the
optimized microstructure of an Fe73sSi13sBgNb3Cul alloy is shown in Fig. 6.4.
This consists of a nanocrystalline bcc Fe-Si phase with an average grain size of
approximately 10 nm (weakly ordered to the 00 3 structure). The grains are
randomly oriented without any preferred orientation. In addition to the grains of
-10 nm, smaller grains can be recognized as indicated by the arrowheads.
These are the fcc-Cu particles that precipitate in the early stage of
crystallization CHonoetal., 1993). YoshizawaandYamauchi (1990) reported
that a combined addition of Nb and Cu is required for producing such a
272 Kazuhiro Hono

nanocrystalline microstructure; thus the role of Nb and Cu for the


nanocrystall ization has been a subject of extensive research. In addition to
these two phases, an amorphous phase remains in the intergranular region as
shown in Fig. 6.4b.

(a)

(b)

Figure 6. 4 (a) Bright field TEM image of the melt-spun Fen5 Si'3.5 B9 Nb 3 CUI alloy
annealed at 823 K for 60 min (FINEMET>. The grains indicated by arrows are fcc-Cu;
(b) HREM image of the same sample. The bee grains are surrounded by the remaining
amorphous matrix.

In order to understand the role of these additives in the nanocrystalline


microstructure evolution, the crystall ization process of an amorphous
Fe73sSi13sBgNb3Cul alloy has been investigated by lDAP (Hono et aI., 1991,
1992) and more recently by 30AP (Hono et al., 1999). According to these
investigations, three phases were identified based on the chemical
compositions in the optimally heat-treated nanocrystall ine microstructure as
shown in Fig. 6. 5 (Hono et al., 1992). The 00 3 phase contains -- 20 -
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 273

25 at. %Si with little Nb and Cu. A small atomic percentage of boron remains
in this phase, and this appears to be a common feature of the ex-Fe particles
crystallized from boron-containing Fe based amorphous alloys (Zhang et ai.,
2001). The remaining amorphous phase is enriched in Band Nb with a small
amount of Si. In addition to these two major phases, a Cu-enriched particle is
observed. This phase is strongly enriched in Cu ( ...... 60% or higher) but still
contains appreciable amounts of the other elements. A separate nanobeam
electron diffraction study revealed that the Cu-enriched particles in the
optimized microstructure were fcc-Cu (Hono et al., 1993). This phase
appears as a grain having a diameter of approximately 5 nm as indicated by
the arrowheads in Fig. 6.4.
100

80
<!) I
t.L. I

o(l 60 I
I

U
::> ,
I

I
40 I
I
I I I
a-FeSi :amo:a-FeSi:
20
(D0 3) II : (D0 3):
I I I
~ I I
0~
0 50 150 200 250 300 350
:!c 30
;z 20
0 ~
.~

C
10
<!)
u
c
0 0 50 100 150 250 300 350
u 30
20
co
10

0 50 100 150 200 250 300 350


40
30
Vi
20
10
0 50 100 150 200 250 300 350
Number of atoms(X50)
Figure 6.5 Atom probe concentration depth profiles obtained from a Fen5 Si ,3 . 58 9 Nb3 Cu,
alloy annealed at 550'C for 60 min (optimal condition).

It was confirmed that the as-melt-spun alloy was a uniform amorphous


phase, and neither structural nor chemical ordering was detected within the
limitation of the spatial resolution of APFIM (atom probe field ion microscope)
and TEM techniques. Here, it should be noted that atom probe data do not
include any information on the chemical short range order (CSRO), even if
274 Kazuhiro Hono

amorphous alloys are compositionally short range ordered as suggested in


literature (Egami, 1983), because APFIM data do not retain the information
on nearest neighbor atoms. However, since atom probe is sensitive in local
compositional change in a nanometer scale, it can detect solute clusters that
are formed even prior to the crystallization reaction. 3DAP elemental maps of
Cu within analyzed volumes of 10 nm x 10 nm x 40 nm in an as-melt-spun
sample and the ones annealed for 5 and 60 min at 400C are shown in Fig. 6.6,
ASQ

5 min at 400C

60 min at 400 C

Figure 6. 6 HREM images and 3DAP Cu maps of as-quenched Fe73.5Si135BgNb3CU,


amorphous alloys and the ones annealed at 400C for 5 and 60 min.

together with their HREM (high resolution electron microscope) images (Hono
et ai., 1999). In the as-melt-spun sample, Cu distribution is uniform,
confirming that it is a chemically homogeneous solid solution. In the sample
annealed for 5 min at 400C, heterogeneous distribution of Cu atoms is
apparent, indicating that Cu clustering occurs. After a 60 min annealing at
400C, Cu clustering is observed more clearly. The number density of the
clusters decreases, while the composition of Cu in the clusters increases. The
HREM image shows that no crystallization occurs up to 60 min at 400C , thus it
was concluded that the clustering observed in this analysis occurs in the
amorphous phase. The number of atoms in each cluster was in the range of 50
to 100 (taking the detection efficiency of the microchannel plate detector into
consideration), and the size of the clusters is approximately 3 nm. The number
density of the Cu clusters estimated from the analyzed volume is in the order of
10 24 m- 3 . The concentration of the Cu clusters has been estimated to be
approximately 12 at. % Cu initially, but it increases as the clusters grow in
size. The structure of these clusters was not identified by HREM, because the
HREM image did not give any fringe contrast corresponding to any crystalline
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 275

structure at these stages. Based on the extended X-ray absorption fine


structure (EXAFS) measurement results, Sakurai et al. (1994) and Ayers
et al. (1998) reported that Cu atoms form clusters having near-fcc symmetry
from the very early stage of the heat treatment. Thus, the Cu clusters
observed in the 3DAP data are believed to have an fcc-like short range
ordered structure, but they do not appear to be distinct fcc-Cu in the initial
stage. Evidence for Cu clustering prior to the onset of the crystallization
reaction in this alloy was also obtained by more recent differential scanning
calorimetry (DSC) and small angle neutron scattering (SANS) results
(Ohnuma et al., 2000).
Earlier studies by Hono et al. (1992), employing the conventional AP
failed to prove that these Cu clusters serve as heterogeneous nucleation sites,
for the Fe-Si primary crystals. Thus, they concluded that the Fe-Si primary
crystals nucleate from the Cu-depleted (Fe-enriched) amorphous matrix as
originally proposed by Yoshizawa and Yamauchi (1990). On the other hand,
Ayers et al. (1998) proposed that the Cu clusters that form in the amorphous
matrix should serve as the heterogeneous nucleation sites for the Fe-Si primary
crystals. The assumption that Cu clusters serve as the heterogeneous
nucleation sites is quite reasonable, but obtaining direct evidence for this
mechanism was very difficult. Using a 3DAP, Hono et al. (1999) succeeded is
showing that Cu precipitates were in direct contact with the Fe-Si particles,
which suggests that the Cu cluster directly serves as heterogeneous nucleation
sites for the Fe-Si primary crystals. The formation of Cu clusters in the
amorphous state can be understood based on the free energy curve of the
liquid phase in the Fe-Cu binary system. As shown in Fig. 6.7a, there is a
tendency of phase separation of Fe-rich and Cu-rich phases in the liquid phase
of Fe-Cu binary alloy (Ohnuma et al., 2001). Thus, the Cu-rich clusters are
expected to form in the liquid state (or in the amorphous phase) in the
FINEMET (Fe-Si-B-Nb-Cu nanocrystalline soft magnetic material) type alloy
prior to the nucleation of the ex-Fe primary crystals.
A number of attempts have been made to improve the magnetic properties
of the Fe735Si135BgNb3Cul alloy by modifying the alloy composition and
substituting some of the alloying elements. Yoshizawa (1999) recently
improved the saturation magnetic flux density, Bs ' of the FINEMET alloy to the
level of 1. 4 T by modifying the original composition to more Fe-rich ones,
i.e., FenSillBgNb3-xCux' Although the optimized Cu content in the original
FINEMET alloy was 1.0 at. %, the one for the modified alloy was found to be
O. 5 at. % as shown in Fig. 6. 8. It is interesting to note that the lower
concentration of Cu gives optimal nanocrystalline microstructure, because it is
usually expected that higher Cu content results in a higher number density of
heterogeneous nucleation sites for ex-Fe primary crystals. Taking the kinetics of
the Cu-clustering into account, Ohnuma et al. ( 2000) explained the optimized
concentration of Cu required to refine the nanocrystalline microstructure in the two
types of FINEMET alloys, i. e., Fe745- x Si 135 BgNb3CU x and Fen Sill BgNb3- xCux .
276 Kazuhiro Hono

'"""'

i~~ i~:~,
"0 ~1000K ~

t
~
~ Fe 20 40 60 ~
SO Cu ~ Fe 20 40 60 SO Cu
eu concentration (at.%)

-Se
Cu concentration (at.%)

!
'3 -12 700 K !=-_ -20
19

700 K

r- 16
~ ~-21
ii -22
~
" -20
e
"- Fe 20
I J

40 60 SO Cu
e -23
"- FeCo 20
J
40 60 SO Cu
Cu concentration (at.%) Cu concentration (at.%)
(a) (b)

Figure 6.7 Free-energy composition curves for liquid phase in (a) Fe-Cu binary alloy and
(b) FeCo-Cu quasibinary alloy systems calculated by the CALPHAD method (Ohnuma
et aI., 2001).

160

140
'"x 120
0

'-'

~ 100
:.c
"'" SO

"0- 60
'.g"
:>
40
"
0::
20
0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
x

Figure 6. 8 Relative permeability of Fen Sill 8 9 Nb3 - x CU x and Fe74.5-x Si t3 . 5 8 9 Nb3 CU x


alloys as functions of x. The samples were annealed at 550'C for 60 min (Ohnuma et al.,
2000).

Using a highly sensitive differential scanning calorimeter (DSC), they


measured the clustering temperature, T clust' and crystall ization temperature,
T x' By measuring the peak shift of TClust and T x as functions of heating rate,
T clust and T x were extrapolated to the heating rate for the industrial anneal ing
condition. In order for the Cu clusters to serve as heterogeneous nucleation
sites for a-Fe, T clust must be lower than T x' i. e., T x> T clust. T clust decreases
as Cu content, x, increases, because the kinetics of the clustering become
faster as the supersaturation of Cu increases, while T x does not show much
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 277

dependence on x. Figure. 6. 9 shows T x and Telust for the two alloy


compositions (Ohnuma et al., 2000). In the original FINEMET composition,
T x is lower than Telust with x = 0.5. In this case, Cu clustering does not occur
prior to the crystallization reaction at 823 K with a heating rate of 20 K/s which

850

800

~
"
~ 750
h'
vi"
-:ro"
~
::l
~ 700
'"0.
E
'"
I-

650
...

600 L...l.-_ _---'--_ _-----'-_ _-----' -'------'-----_--'-

0.001 0.01 0.1 I 10 100


Heating rates, f3 (K/s)
(a) Fe74.5_xSit3s89Nb3Cux
850
I
---------------------~-
800

650

600 CL-_--'--.....L_ _---"-_ _---"-_ _---L---"-_-'-


0.001 0.01 0.1 1 10 100
Heating rates, f3 (K/s)
(b) FenSi1l89Nb3_xCUx

Figure 6.9 Cu clustering temperature, Telus" and crystallization temperature, T x ' of (a)
Fe74.5-xSi13589Nb3CUx(X=0, 0.5,1.0,1.5) and (b) FenSitlB9Nb3-xCUx(x=0, 0.6,
1.0) alloys plotted as functions of heating rate. The plots are extrapolated based on the
Kissinger equation (Ohnuma et al., 2000).
278 Kazuhiro Hono

corresponds to the industrial annealing condition, thus the crystal grain size
=
cannot be refined. For x 1.5, clustering occurs at a much lower temperature
than T x' thus the number density of Cu clusters decreases by coarsening
=
before the crystallization occurs at T x' In the alloy with x 1.0, Cu clustering
occurs just before the crystallization event, thus the highest number density of
Cu clusters can serve as the heterogeneous nucleation sites for ex-Fe; hence,
the largest number of ex-Fe nucleate in this alloy composition. On the other
hand, in the modified FINEMET composition, T clust is just below T x at x = 0.6,
but T clust is too low at x = 1.0; hence the optimal x for this alloy composition is
0.6. This work nicely demonstrated that the interplay of the kinetics of Cu
clustering and the primary crystallization is very important to control the
nanocrystalline microstructures in Cu-containing Fe-based soft magnetic alloys.
In an attempt to decrease the coercivity, Tate et al. (1998) substituted
Fe for AI based on the fact that FeAISi has lower magnetocrystalline anisotropy
constant than FeSi alloy. They found that the substitution of 2 at. % AI for Fe
significantly reduces the dc coercivity, while saturation magnetostriction
increases. A 3DAP study by Warren et al. (1999) showed that AI is
partitioned into the Fe-Si phase forming Fe-19Si-3AI solid solution. Although
the inherent magnetocrystalline anisotropy decrease by substitution of AI for
Fe, saturation magnetostriction cannot be optimized because As of the Fe-
19Si-3AI phase becomes positive, while that for Fe-20Si phase is negative.
The overall magnetostriction constant As is a balance between those for
crystall ine and amorphous phases, i. e. ,

Since A=mo is positive ( ....... 20 x 10- 6), A~ should be negative to balance the
As' Thus, the substitution of AI for Fe in Fe-20Si nanocrystal did not reduce
the magnetocrystalline constant. As seen from this example, determining
solute partitioning in nanocrystalline alloy is important to understand their
magnetic properties. For this purpose, the atom probe is the most suitable
experimental technique.
In order to apply a good induction anisotropy to the FINE MET type alloys
for various high frequency applications, Yoshizawa (2001) replaced part of Fe
with Co. Although saturation magnetization M s and quality factor, Q, were
significantly improved by replacing Fe with Co up to 15 at. %, the coercivity
increased significantly when Co composition exceeded 20 at. %. Figure 6. 10
shows 3DAP analysis results of Fe788-xCoxSig8gNb24Cuo6 (x = 5, 20, 40)
alloys with three different Co contents. It can be clearly seen that the number
density of the Cu clusters decreases as Co content in the alloy increases.
When x = 60, the number density of Cu clusters became too scarce to be
detected using the atom probe technique, whose analysis volume is typically
10 nm x 10 nm x 100 nm. This result suggests that the driving force for Cu
clustering decreases when Fe is replaced with Co, thereby decreasing the
number density of the heterogeneous nucleation sites for the ex-Fe primary
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 279

crystals. Because of this, the grain size of the FINEMET alloy increases when
Fe is replaced with Co, thereby He increases. The driving force for the
clustering of Cu in the Fe-based amorphous alloy is its large positive enthalpy
of mixing. For Cu clusters to serve as heterogeneous nucleation sites for bcc-
Fe, the phase separation must occur in the amorphous phase. In fact, this was
confirmed by the 3DAP analysis result shown in Fig. 6. 6. Since the enthalpy
of mixing is lower in a positive value between Co and Cu, it is speculated that
the replacement of Fe with Co would reduce the driving force for phase
separation of Cu-enriched phase in the amorphous. In fact, the free-energy
composition curve for the liquid phase of FeCo-Cu quasibinary apply shown in
Fig. 6. 7b indicates that the driving force for phase separation between Fe-rich
amorphous and Cu-rich amorphous decreases significantly compared to that for
the Fe-Cu binary alloy. Thus, in the Co-containing Fe-based amorphous
alloys, addition of Cu is not useful to refine the crystal grain size.
Fe 78.8_.,CoxSi9B9Nb24CU06

x=5 ,< :; ~.:~: ;~::j;: :::ti :;.:~~:;:;;,:~:\.~:;~':::T~:/.:.:~:\:-~

-50 nm
Figure 6.10 3DAP Cu maps of Fe788-xCoxSigBgNb2,CU06 alloys annealed for 10 min at
550'C for 10 min.

6.3.2 NANOPERM (Fe -Zr-B( -Cll) )

The Fe-Zr-B nanocrystalline alloy was originally developed by Suzuki et al.


(1990). Thereafter, a series of Fe-TM-B (transition metal TM=Zr, Hf, Nb)
type nanocrystalline soft alloys was developed as NANOPERM (Makino et aI.,
1997). The crystallized microstructure consists of a large portion of
nanocrystalline ex-Fe particles embedded in a small fraction of the residual
amorphous phase. The thermal analysis and X-ray diffraction results (Suzuki
et al., 1991> showed that the crystallization of the amorphous Fe91Zr7B2 alloy
occurs through two distinct stages, i. e. ,
amorphous-ex-Fe + amorphous' -ex-Fe + Fe2 Zr + Fe3 Zr.
The first stage is the primary crystallization of the ex-Fe phase from the
amorphous matrix, and the second stage is the eutectic crystallization of the
280 Kazuhiro Hono

remaining amorphous phase. The mechanism of the nanocrystalline


microstructure evolution seems to be different from that in Fe-Si-B-Nb-Cu
alloys, because the Fe-Zr-B alloy does not require the addition of a nucleation
agent such as Cu with a large positive enthalpy of mixing with Fe. Hence, it is
expected that there is a large density of nucleation sites or that the
homogeneous nucleation rate is high in the Fe-Zr-B alloys.
Zhang et al. (1996a) made a comprehensive atom probe study on the
nanocrystallization process of the FegoZr7B3 amorphous alloys. By APFIM and
TEM, the as-quenched FegoZr7B3 alloy was confirmed to be a homogeneous
supersaturated solid solution with the amorphous structure. As possible nuclei
for the primary crystals, some regions that display a contrast which is similar
to that of the medium range ordered (MRO) structure proposed by Hirotsu
et al. (Hirotsu et aI., 1986; Nakamura et aI., 1994) were observed by
HREM. This feature was observed even in the as-quenched alloy, but after
anneal ing below the crystall ization temperature, such MRO-I ike contrast was
observed more clearly. This indicates that very small MRO domains are
developed during annealing and that these may act as the nucleation sites for
the primary crystals. Suzuki et al. (1994) studied the kinetics of the primary
crystallization of the Fego Zr7 B3 amorphous alloy and found that the Johnson-
Mehl-Avrami exponent ranges from 1 to 2. 1 depending on the temperature.
This suggests that the crystallization occurs mostly on site-saturated quenched-
in nuclei. Thus, if we assume that the nucleation of a-Fe occurs from the MRO-
like feature observed in the as-quenched amorphous alloy, the value of the
Avrami exponent can be reasonably interpreted. The concentration profile of
Zr directly measured by the atom probe showed the depletion of Fe and
enrichment of Zr atoms at the aj amorphous interface, which suggests that the
metastable local equilibrium is achieved at the interface of the growing primary
crystal and the particle growth is controlled by the diffusion of Zr.
The optimized microstructure of this alloy is also composed of a-Fe and
the remaining amorphous phase as shown Fig. 6. 11. The amorphous phase
does not necessarily remain at all grain boundaries as can be seen from the
sharp grain boundary indicated as A-A in Fig. 6.12. The volume fraction of the
a-Fe in this alloy appears to be much higher than that observed in the
FINEMET alloy, and this is the reason why the saturation magnetization of this
alloy is higher than that of FINEMET. Atom probe concentration depth profiles
of the same sample show that there are Fe-enriched and Zr-enriched phases as
shown in Fig. 6. 12. The Zr concentration in the Zr-enriched phase is almost
equal to that expected from the Fe3 Zr phase (25 at. % Zr), but no diffraction
evidence for the presence of equilibrium Fe3 Zr was observed in this stage.
Hence, this Zr enriched phase is believed to be still amorphous as observed in
Fig. 6. 11. In these phases, B concentration is as high as 10 at. %, but an
appreciable amount of boron still remains in the a-Fe phase. On the other
hand, Zr is almost completely rejected form the a-Fe phase. It should be noted
that neither Zr nor B is segregated at the ajamorphous interfaces in this stage,
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 281

Figure 6. 11 HREM image of the nanocrystalline Fego Zr7 8 3 alloy after annealing for 5 min
at 923 K from the amorpyhous state.

100 I 200: 300: 500 600


I I

~ 20
~

~ 10
"g
u
o
co
100 200 300 400 500 600
Number of atoms (x 50)

Figure 6. 12 Atom probe concentration depth profiles of the Fego Zr7 8 3 melt-spun alloy
annealed at 923 K for 5 min.
282 Kazuhiro Hono

suggesting that partitioning of Fe and B are almost completed.


Unlike FINEMET type alloys, the Fe-Zr-B amorphous alloy forms
nanocrystall ine microstructure without the addition of Cu. This is probably
because there are quenched-in nuclei in the as-quenched amorphous alloy and
the nucleation of the primary ex-Fe crystals occurs from these quenched-in
nuclei as described above. However, adding a small amount of Cu to the
FegoZr7B3 amorphous alloy further refines the grain size as shown in Fig. 6.13
(Suzuki et al., 1991). Of particular interest in these micrographs is that the
orientation of the ex-Fe nanocrystals in the crystallized microstructure of the
Fes6Zr7B6Cui alloy was entirely random, while that of the Fes7Zr7B6 alloy
showed aggregates of several grains with the same orientation. This indicates
that the addition of Cu enhances uniform distribution of randomly oriented ex-Fe
particles that are heterogeneously nucleated from a large number density of
fcc-Cu, while heterogeneous nucleation of ex-Fe occurs from the ex-Fe/
amorphous interfaces of the ex-Fe particles that nucleate earlier in the Cu-free
sample.

(a) (b)

Figure 6. 13 Bright-field and {all} dark-field TEM images of (a) FeB7Zr7B6 and (b)
FeB6 Zr7 B6Cu, alloys annealed for 1 h at 600C (Suzuki et aI., 1991).

In order to investigate whether or not Cu clustering occurs in this system


like in the FINEMET alloys, Zhang et al. (1996b) observed solute distribution
in the course of the crystallization reaction by APFIM. They confirmed that all
alloying elements were homogeneously dissolved as an amorphous
supersaturated solid solution in an as-quenched FesgZr7B3Cul alloy. After
annealing at 673 K for 60 min, it was confirmed that the specimen was still
amorphous, but atom probe concentration depth profiles clearly showed that
Cu atoms form clusters. More recent work using 3DAP and HREM by Ohkubo
et al. (2001) has convincingly shown that the fcc-Cu nanoparticles that
precipitate prior to the crystallization provide heterogeneous nucleation sites
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 283

for the a-Fe primary crystals as shown in Fig. 6. 14. The Cu atom map clearly
shows that Cu atoms form clusters. Based on the concentration of Fe, the
shape of an a-Fe primary crystal can be displayed with an isoconcentration
surface. It is clearly seen that the a-Fe primary crystal is in direct contact with
one of the Cu clusters. The reason why Cu clusters trigger the nucleation of
a-Fe was thought to be due to the coherent interface that fcc-Cu provides for
a-Fe with orientation relationships (OR) of (111)tcc//(011)bcc' which was
directly shown by Ohkubo et al. (2001) by using HREM and the nanobeam
diffraction method. Figure 6. 15a shows that the a-Fe and Cu particles have
Kurdjumov-Sachs OR, i. e., (011)Fe//(111)cu and [111Je//[101]cu' In
addition, Nishiyama-Wasserman OR was also confirmed, i. e., (110 )Fe/ /
(ll1)cu and [OOl]Fe//[Ol1]cu' Another similar OR described as (llO)Fe//
(ll1)cu' [113]Fe//[101]cu was also found as shown in Fig. 6.15b. All these
OR's appear as a result of the similar atomic configuration of {11 O} Fe and
{ 111 }Cu' Since {111} cu can provide low interfacial energy for {11 0 }Fe' the
activation energy for heterogeneous nucleation from the fcc-Cu particles
becomes lower than that for the homogeneous nucleation.

Figure 6. 14 3DAP elemental map of Cu atoms in Fes9 Zr7 8 3 CU1 amorphous alloy annealed
at 733 K for 60 min. The position of an ex-Fe primary crystal is indicated with an
isoconcentration surface. This clearly shows that nucleation of ex-Fe crystal occurs in direct
contact with Cu clusters.

In Fe-Zr-B nanocrystalline amorphous alloys, many attempts were made


to improve the soft magnetic properties by modifying the alloy compositions,
substituting alloying elements, and adding quaternary and quinary elements.
284 Kazuhiro Hono

(a) (b)

Figure 6. 15 HREM image of Cu and Fe nanoparticles and their nanobeam diffraction


patterns observed in Fes9 Zr7 8 3 CUI amorphous alloy annealed at 733 K for 60 min.

Inoue et al. (1996) reported that addition of Si improves the permeability of


Fe-Zr-B alloys as a result of zero saturation magnetostriction. They found that
addition of 4 at. % Si makes the saturation magnetostriction zero. In
nanocrystalline Fen5 Si 13 . 5B9 Nb 3 CUI alloy (FINEMET), Si partitions into the
ex-Fe phase up to approximately 20 at. % (Hono et aI., 1992). Hence, it was
also anticipated that Si would partition into the ex-Fe phase in the
nanocrystalline Fes7Zr7 Si 4B2 alloy. However, direct atom probe measurement
results showed that Si was preferentially partitioned in the remaining
amorphous phase in an Fes7Zr7Si4B2 alloy annealed at 873 K for 60 min. (Zhang
et al., 1996c). Si is rejected from the ex-Fe particle during the nucleation-and-
growth process from the amorphous matrix, and enriched in the residual
amorphous phase. The enrichment of Si in the residual amorphous phase is
surprising, because in FINEMET type alloy, Si does partition in the ex-Fe
phase. This reverse partitioning behavior was explained in terms of a strong
interaction between Si and Zr atoms compared to that between Si and Fe. The
inverse partitioning behavior of Si in Fe-Zr-B nanocrystalline soft magnetic
alloy explains the dependence of the magnetostriction constant on Si contents
as shown in Fig. 6. 16. Whereas the saturation magnetostriction (As) of
Fen5SixB235-xNb3Cu, nanocrystalline alloy decreases from positive to negative
values as a function of Si content (Herzer, 1989), As of Feso- x Zr7 Six ~
nanocrystalline alloy shows the opposite dependence on the Si content. In the
case of Fen5SixB235-xNb3Cu, nanocrystalline alloy, Si is partitioned in the ex-
Fe phase, forming Fe-20Si bcc solid solution. The net magnetostriction
constant is determined by the balance between magnetostrictions of ex-Fe and
amorphous matrix phases, i. e.,
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 285

40
o Fe90-xZr7SixB3 amorphous 0 Fe73 sSi xB2) s-xNb3CUI amorphous
35
Fe90-xZr7SixB) annealed Fe7JsSixB2Js-xNb)Cul annealed
30
25 o 0 rL

1:' 20
x 15
~ 10

~b;:-----.......I..==!::::::::~==~S~~
-5 '=----_---'---_----.lL-_--L-_-----'-_ _-'---_----'--_ _L--_---'--_-----'
o 2 4 6 8 10 12 14 16 18
Si concentration (at.%)

Figure 6. 16 Magnetostriction constant of Fe90-xZr7SixB3 and Fe73.5Six B23.5-x Nb3CUI


amorphous and nanocrystalline alloys as a function of Si content. Open circles show the
magnetostriction of as-quenched Fe90- x Zr7 Six B3 amorphous alloy, and closed circles show
the magnetostriction of Fe90- x Zr7 Six B3 alloy annealed at 873 K for 60 min (after Inoue
et al., 1996) . Open squares show the magnetostriction of as-quenched
Fe735 Six B235 - x Nb3Cu, amorphous alloy, and closed squares show the magnetostriction of
Fe735SixB23.5-xNb3CU, alloy annealed at 813 K (after Herzer, 1989). The figure is
adapted from (Zhang et aI., 1996b).

In the case of the Fe-Si-B-Nb-Cu alloy, volume fraction of ex-Fe, Va' increases
as a function of Si content, because Si partitions into the ex-Fe phase forming
Fe-20Si solid solution. As Fe-20Si has high negative magnetostriction ('" - 20
X 10- 6 ), the net magnetostriction As would reduce as a function of volume
fraction of Va' When Fe-Si-B-Nb-Cu alloy with Si less than 16 at. %is
crystallized from the amorphous phase, As is still positive, because Va of
Fe-20Si is not high enough to make the total magnetostriction negative. As Si
content of the alloy increases, Va increases and the negative factor from the
Fe-20Si phase increases; then, the net magnetostriction becomes negative.
On the other hand, the Fe-Zr-B alloy without Si shows negative
magnetostriction on crystallization, because the primary particle of ex-Fe has
negative magnetostriction ('" - 4.5 X 10- 6 ) . Since Si is rejected from the
ex-Fe and enriched into the amorphous matrix phase, it is expected that Va of
Fe-Zr-Si-B reduces as a function of Si content. As Va decreases, the net
magnetostriction becomes positive by passing zero magnetostriction at 4 at. %
Si. Hence, the opposite tendency of magnetostriction change as a function of
Si contents in Fe-Si-B-Nb-Cu and Fe-Zr-B-Si alloys can be explained by the
opposite partitioning tendency of Si. As can be seen from this example,
experimental measurement of solute partitioning in the nanocrystalline
magnetic materials is essential to understand the magnetic properties. The
atom probe technique is the only experimental technique that is capable of
286 Kazuhiro Hono

directly determining solute partitioning in nanocrystalline phases embedded in


the matrix phase.
Surface crystallization is commonly observed in melt-spun ribbons with
relatively low glass forming ability during rapid solidification. It also occurs
during crystall ization from fully amorphous ribbons. Since the surface
crystallization tends to cause a strongly textured layer of gross grains, it is not
desirable for soft magnetic materials. The softness of the nanocrystalline
alloys is due to the averaging effect of the magnetocrystalline anisotropy;
hence, having crystallographically textured microstructure results in
degradation of soft magnetic properties, i. e., increase in He (coercivity) and
decrease in lJ(magnetic moment). Recently, Makino et al. (2000) reported that
saturation magnetostriction, As' can be controlled by mixing the Fe90 Zr7 B3 alloy
with negative magnetostriction and the Fe84 Nb 7 B9 alloy with positive
magnetostriction in various ratios. In a series of nanocrystalline
(Fe90 Zr7 B3) 1- x (Fe84 Nb 7B9) x alloys, they found that the zero magnetostriction
was obtained when x = O. 8, but the permeability, lJe' of the alloy was not
improved as expected from the zero magnetostriction; on the contrary, lJe is
=
degraded at x 0.8. By careful TEM observation of the (Fe90 Zr7 B3)02 (Fe84
Nb 7B9 ) 08 alloy with zero magnetostriction, Wu et al. ( 2001) found that
strongly textured surface crystallization occurs in the alloy with a poor
permeability as shown in Fig. 6. 17, while relatively uniform nanocrystalline

(a) (b) (c)

Figure 6. 17 TEM bright field image of (a) free-surface, (b) inner part, and (c) wheel-
contacted surface of an optimally heat treated Fe85Zr12Nb5SBs alloy melt-spun-ribbon.

microstructure develops in the alloy with high permeability. Surface


crystallization was more pronounced in the free surface than in the wheel
contacted surface as shown in Fig. 6. 17, and the gross grains were all
oriented to the {OO 1} planes. By removing these surface crystallized layers
from the (Fe90Zr7B3)08(Fe84Nb7B9)o2 alloy ribbon by mechanical grinding, it
was shown that the permeability was increased. This observation convincingly
showed that the degraded soft magnetic property was due to the formation of
the surface crystallized layers. The detailed mechanism of such surface
crystallization is not understood very well, but they found that the surface
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 287

crystallization can be suppressed by inducing uniform crystallization throughout


the ribbon sample, i. e., by adding a small amount of Cu. This is because
uniform nanocrystalline microstructure was developed in the entire volume of
the sample as a result of uniform distribution of the heterogeneous nucleation
sites for the ex-Fe primary crystals. By removing the surface-crystallized layer,
the highest permeability was obtained from the alloy composition of zero
magnetostr iction.

6.3.3 HITPERM CCFe,Co)-Zr-B-Cu)

In order to increase the saturation magnetic flux density, B s ' or to increase


the Curie temperature, the natural choice as an alloying element is Co. By
replacing Fe with Co up to 30 at. %, the saturation magnetization increases as
Slater-Pauling curve indicates. Another benefit of Co addition to the Fe-based
alloy is the increase of the Curie temperature. Willard et al. (1998, 1999)
developed a derivative nanocrystalline soft magnetic material of NANOPERM
for high temperature applications by replacing Fe of FessZr7B4Cui with Co to a
composition of Fe44 C0 44 -Zr7 B4CUl , HITPERM. The crystallized microstructure
is composed of nanoscale ex' -FeCo particles (B2 structure) embedded in a
residual amorphous phase. The ac permeability has been found to maintain a
high value of 1800 up to a frequency of -- 2 kHz. The core loss in room
temperature is competitive with that of commercial high temperature alloys and
the high magnetization persists to the ex-y phase transformation temperature
at 980'C .
In the HITPERM alloy, Cu was added as a nucleation agent from the
beginning of the alloy design (Willard et al., 1998), and thus whether or not
Cu was really required for the formation of nanocrystalline microstructure has
not been tested. As shown in Fig. 6. 8b, the driving force for the formation of
Cu clusters decreases when Fe is replaced with Co. Hence, it is worthwhile
examining whether or not Cu addition is required for obtaining optimum soft
magnetic properties in the HITPERM alloy. In view of this, Ping et al. (2001a)
studied the clustering and/or partitioning behavior of Cu atoms during the
crystallization process of an amorphous Fe44C044Zr7B4Cui alloy using the 3DAP
technique. Figure 6. 18 shows 3DAP maps of Zr and Cu atoms in the Fe44 Cow
Zr7 B4CUl nanocrystalline alloy annealed at 550'C for 1 h. This shows that there
are Zr-enriched and Zr-depleted phases, the former is the remaining
amorphous phase and the latter is the ex' -FeCo phase. Unlike the case of
FeS9Zr7B3Cul amorphous/nanocrystalline alloys, there is no evidence for Cu
clusters. Cu atoms are partitioned in the remaining amorphous phase without
any indication of clustering or segregation. This result indicates that Cu
clustering does not occur in the alloy in which a substantial amount of Co is
replaced with Fe, as was observed in the Co-containing FINEMET alloy.
Although the composition of the FeCo phase was designed to be Fe50 C0 50 , the
288 Kazuhiro Hono

atom probe result showed that the composition of the ex' was approximately
Fe6oC04o, This is because Co preferentially partitions in the remaining
amorphous phase, because Co is more attracted to Zr than to Fe. The
enthalpy of mixing for Co and Zr is very negative, - 197 kJ/mol, while that for
Fe and Zr is - 118 kJ/ mol. This suggests that Co and Zr are more attractive.
Because of this, Co is weakly enriched in the remaining phase rather than
enriched in the ex' -(Fe, Co) solid solution as shown in Fig. 6. 18b.

Cu

-40 nm
(a)
~ I

z
" I 1.1 at.%
S:60 1
N ~

~ 400 amorphous
]30~
" 300
.S: 0 200 400 I 600 800
<l) ;z.
u.
"-
0 200 ~

....
<l) "s: 30
.n 100
E co
::l "-
0
15
;z.
0 200 400 600 800 OJ
.n
~ 400 E
::l
o 200 400 600 800
"s: 300
;z.
0
u ~

"-
0 200 " 10
.S:
....<l) ::l
.n U 0.5 at.%
100 5
E "-
0
::l
;z. OJ
0 200 400 600 800 .n
E
o 200 400 600 800
Total number of detected ions ::l
;z. Total number of detected ions

(b)

Figure 6. 18 3DAP results of the Fe44 0044 Zr7 B4Ou, nanocrystalline alloy (HITPERM)
annealed at 550'0 for 60 min. (a) Zr and Ou maps, and (b) integral concentration depth
profiles of Fe, 00, Zr, B, and Ou.

Figure 6. 19 compares the nanocrystalline microstructure of Cu-containing


and Cu-free HITPERM alloys. Although the grain size itself is extremely fine
(-10 nm), aggregates of several grains with the same crystallographic orientation can
be seen in both alloys. This is in contrast to the case of Fe-Zr-B alloy, where the Cu-
containing alloy showed rather uniform random distribution of nanograins, while
Cu-free alloy form aggregates of the grains with the same orientation as shown
in Fig. 6. 14 (Suzuki, 1991). This suggests that the Cu clusters that serve as
heterogeneous nucleation sites for bee-Fe are not present in the HITPERM
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 289

alloys regardless of the presence of Cu. Based on these microstructure


investigations, it was concluded that Cu is not a required alloying element in
the HITPERM alloy, and in fact, similar soft magnetic properties are obtained
regardless of the presence of Cu in the alloy.

(a) (b)

Figure 6. 19 Bright field TEM images and SAED patterns of (a) Fe" Co" Zr7 B4CUI and
(b) Fe44.sC044.sZr7B4 alloys annealed at 550'C for 60 min.

6.4 Microstructural Evolution in Nanocomposite


Magnets

Nanocrystallization of an amorphous phase is also useful to produce


nanocomposite hard magnetic materials. Nanocomposite magnets are
composed of nanoscale soft and hard magnetic grains, and due to their
exchange coupling, they show a higher remanence that exceeds the theoretical
limit for isotropic magnet (0. 5M s )' Because of this remanence enhancement
effect, relatively high maximum energy product, (BH) max' can be obtained
from the alloy compositions containing lower amounts of rare earth than is
required for the single phase Nd2 Fe14 B magnet. This makes the rare earth
nanocomposite magnets economically attractive. Although there are several
combinations of hard and soft phases to produce nanocomposite magnets,
most extensively studied nanocomposite hard magnetic materials are based on
Fe-Nd-B ternary alloy. Depending on the alloy compositions, Fe3 B/Nd2 Fe14 B,
a-Fe/Nd 2 Fe 14B and their mixture are possible. Since enhanced remanence is
due to the exchange interaction between the hard and soft grains, reduction of
the grain size less than the domain wall width of the hard phase is necessary.
However, if the grain size becomes too small, magnetocrystalline anisotropy
of the hard grains is averaged out due to the exchange coupling between the
hard phases, resulting in a decrease of coercivity. Based on the
290 Kazuhiro Hono

micromagnetic modeling, maximum coercivity and (BH)max are expected from


the grain size of approximately 20 nm (Fischer et al., 1996). However,
obtaining such a fine scale nanocomposite microstructure is not easy. Hence,
various attempts were made to refine the grain size of the nanocomposite
magnets by microalloying and process control. In this section, examples that
the atom probe technique clarified the roles of microalloying elements in
improving hard magnetic properties are overviewed.

The Fe3 B/Ndz Fe14 B nanocomposite was first reported by Coehoon et al.
(1988) by rapidly solidifying and crystallizing the Ndu Fen.5 B 18 .5 alloy. The
crystall ization progresses in two stages:
amorphous-Fe3 B + amorphous-Fe3 B + NdzFe14 B.
When annealing temperature is lower than 660'C or when heating rate is low,
an intermediate Ndz Fe23 B3 phase forms, which decomposes into Fe3 Band
NdzFe 14B 1 phases at above 660'C .
Although this nanocomposite showed very large remanence and an
acceptable level of (BH) max for a low cost material, the coercivity was not
high enough for many applications. Hence, many alloying elements were
tested to improve the magnetic properties, in particular, to increase the
coercivity (Hirosawa et ai., 1993; Kanekiyo et ai., 1994; Archambault and
Pere, 1999; Bernardi et al., 2000). There are several ways to improve the
hard magnetic properties of nanocomposite magnet materials. Heavy rare-
earth elements such as Dy and Tb increase the intrinsic coercivity of the hard
phase by substituting for Nd. However, these elements usually do not have
much influence on the microstructures. Co is commonly alloyed in the Fe3 B/
NdzFe14 B nanocomposite, because it increases the coercivity and the Curie
temperature. However, Co does not show strong partitioning behavior and,
hence it does not give much influence on the microstructure either. Cr addition
significantly increases the coercivity of the Fe3 B/NdzFe14 B nanocomposite at
the expense of remanence (Hirosawa and Kanekiyo, 1996). Thus, by
changing the Cr content x in the Ndu Fen- x B 18 .5Cr x alloy, it is possible to
produce a series of nanocomposite magnets with high coercivity and low
remanence or with low coercivity and high remanence depending on the
performance required for various applications (Hirosawa et al., 1999). As the
reason for the coercivity increase with the addition of Cr, Uehara et al.
(1998) reported that the addition of Cr stabilizes the Fe3 B phase, by which
the formation of the NdzFeZ3 B3 phase is suppressed and the volume fraction of
the NdzFe14 B phase increases. However, it should be noted that the
Fe3B/NdzFe14B nanocomposite can be formed when the NduFenBl8.5 ternary
alloy is isothermally annealed above 660'C with a sufficiently high heating rate
(Ping et al.. 1999). In this case. the coercivity obtained from the Cr-
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 291

containing alloy is still higher (Wu et ai., 2001). Thus, the underlying
mechanism of the coercivity increase by the addition of Cr may be due to the
Cr partitioning in the Fe3 B soft phase, which would reduce the exchange
coupling between the hard grains (Hirosawa and Kanekiyo, 1996). The grain
size dependence on the Cr concentration has not been examined, so there is
another possibility that the coercivity change may be related to the
microstructural change. The mechanism of the coercivity increase the addition
of Cr needs to be more thoroughly investigated.
Ga is also known to improve both remanence and coercivity (Hirosawa
et al., 1993). Although this was attributed to the grain size refinement
(Hirosawa et ai., 1993), these elements do not have a noticeable effect of
grain size refinement as shown in Fig. 6. 20. The average grain size of Ga-

(a) (b)

(c) (d)

Figure 6.20 TEM bright field images of (a) Ndu Fen 8'85' (b) Ndu Fe7' 8'8.5 C03 , (c)
Ndu Fe76 8'85 Ga, and (d) Ndu Fe73 8'8 5C03Ga, alloys annealed at 973 K for 10 min.

containing alloy appears to be larger than that of the Ga-free alloys. Co


addition does not affect the grain size much. Thus, the mechanism of the
magnetic property improvement by the addition of Ga may not be due to the
grain size refinement by the microalloying of Ga. As Ga partitions in the
Nd2 Fe14 B phase with a weak tendency to segregate at the Fe3 B/Nd2 Fe14 B
interface (Ping et al., 1998), this may weaken the exchange interaction
between hard grains like Cr partitioning to the Fe3B phase. This may increase
the coercivity by suppressing the averaging effect of magnetocrystalline
292 Kazuhiro Hono

anisotropy of the hard phase. Due to the partitioning of Ga in the Nd2Fe14 8


phase, the volume fraction of the hard phase may also change.
Ping et al. (1999) attempted to observe the effect of grain size difference
on the magnetic properties by changing the grain size, keeping the other
factors as same as possible. For this purpose, they chose Cu for refining the
microstructure, as Cu was known to be very effective in refining the grain size
of the nanocrystalline soft magnetic materials as described above. In addition,
they examined the effect of Nb in conjunction with the Cu effect. Figure 6.21 a
show bright field TEM micrographs of NduFen818.S' Ndu Fe76.8 8 18sCUO2 ,
NduFe7S.8818.sNblCUO.2 and NduFe76818.sNb1 alloys annealed at 660'C for 10
min, respectively. The average grain size of the ternary alloy is approximately
30 nm, while those of the Cu-containing quaternary and Cu-and Nb-containing
quinary alloy is approximately 17 nm and 12 nm, respectively. On the other
hand, the grain size of the Nb-containing quaternary alloy ('" 50 nm) is larger
than that of the ternary alloy. It should be noted that adding only O. 2 atomic
percent of Cu to the ternary alloy is effective to reduce the grain size of the
optimum nanocomposite microstructure. The microstructure is composed of
Fe38 and Nd2Fe14 8 grains in the ternary and quaternary alloys, and there is no
change in the constituent phases in these two alloys. The demagnetization
curves for the NduFen818.S' NduFe768818SCUO.2 and NduFe7S.8818.sNblCU02
melt-spun ribbons (Fig. 6. 21 b) show that the coercivity and (BH) max of the
quaternary and quinary alloys are significantly improved compared with those
of the ternary alloy. The improvements in the magnetic properties in this case
is attributed to the grain size refinement.
In order to understand the reason why the addition of such a small amount
of Cu (0.2 at. %) is so effective in the grain size refinement, the solute
distribution in the early crystallization stage has been examined by a 3DAP.
Figure 6. 22a shows a TEM bright field micrograph of the alloy annealed at
530'C for 30 min. The crystalline particles observed in this micrograph are all
Fe38 embedded in the amorphous matrix, and the second stage crystallization
of the Nd2Fe14 8 phase does not occur yet in this stage. Figure 6. 22b shows a
3DAP elemental map of Cu and Nd in an analysis volume of approximately
'" 17 nm x 17 nm x 56 nm, where small and large black spots correspond to Nd
and Cu atoms, respectively. Fe38 particles can be easily recognized based on
the atomic distribution of Nd because the Nd atoms are fully rejected from the
Fe38 phase. From Fig. 6. 22b, it can be clearly seen that each Cu cluster is
located in direct contact with the Fe38 particles. This result suggests that
Cu-enriched regions (or clusters) serve as heterogeneous nucleation sites for
the Fe38 primary crystals. More detailed 3DAP analysis results have shown
that Nd is also enriched in the Cu enriched cluster, but the total atomic fraction
of Cu and Nd in the cluster was less than 15%. Thus, they are still believed to
be amorphous. So it was proposed that the nucleation of the Fe38 primary
crystals was triggered by a chemical factor. Since Nb and Cu are enriched in
the clusters, boron atoms would pile up at the Cui amorphous interface when
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 293

(a)

10 min at 660C
.... .~

- - Nd 4sFenBlss
............ Nd 4SFe76 sB ls sCUo 2
- - - Nd4.sFe7S.sBI8 sNb ICU O.2

-0.4 -0.3 -0.2 -0.\ 0 0.1 0.2 0.3 0.4


(b)

Figure 6.21 (a) TEM bright field micrographs of Nd45 Fen B'8.S ,Nd45 Fe768 B 18S CU02 and
Nd45Fe7s8B,8sNb,Cuo2 alloys anneaied at 933 K for 10 min. The average grain size is
about 30,17,12 and 43 nm, respectively; (b) H cJ ' B, and (BH)ma> of Nd45FenB'8.s,
Nd45Fe76.8B,8.sCuo.2' Nd45Fe75.8B,8sNb,Cuo2 melt-spun ribbons annealed for 10 min at
various temperatures.
294 Kazuhiro Hono

Cu clusters form. Since the Fe concentration of the alloy (76.8 at. %) is very
close to that in the Fe3 B phase, Nd depletion and B enrichment would be
sufficient to form the local composition of Fe3 B. Since Nd has strong affinity for
Cu (large negative enthalpy of mixing), when Cu atoms aggregate, Nd atoms
are attracted to the Cu atoms, resulting in co-segregation of Cu and Nd. As a
result, enrichment of B and depletion of Nd occur at the Cui amorphous
interface, which benefit the formation of the Fe3 B phase adjacent to the Cu
clusters.

(a) (b)

Figure 6. 22 (a) TEM micrograph and (b) 3DAP map of Cu and Nd atoms of the
Nd45Fe76.8B185Cuo.2 alloy annealed at 530'C for 30 min. This is in the primary
crystall ization stage of Fe3 B from the amorphous matrix.

Combined addition of Cu and Nb further reduce the grain size. 3DAP


analysis results showed that the Cu clusters that form in the early
crystallization stage trigger the nucleation of the Fe3 B primary particles in the
Ndu Fe758 B I85 Nb 1CUO.2 alloy as well. Nb is rejected from the Fe3 B primary
particle and partitioned in the remaining amorphous phase. When the
remaining amorphous phase crystallizes, Nd2Fel4 Band Fe23 B6 phases are
formed in the quinary alloy, and Nb is exclusively partitioned in the Fe23 B6
phase. The Fe23 B6 phase is not observed in the ternary and quaternary alloy,
hence the Nb addition induces the formation of the Fe23 B6 phase in the second
stage of the crystallization.
Although Nb and Zr are similar atoms, their microalloying effects on
Fe3 B/Nd2Fe14 B are different. The addition of only 0.2 at. % of Zr is effective
in refining the microstructure (Kaj iwara et al., 2001). Zr is preferentially
partitioned in the Nd2Fe14 B phase, and evidence for segregation of Zr at the
Fe3 B/Nd2Fe14 B interface was obtained by recent 3DAP analysis. Since Zr
does not have much solubility in both Fe3 Band Nd2Fe14 B phases, when the
amount of alloying is increased, it may form some additional phase. However,
since Zr is effective in refining the microstructure with a small amount, it is a
good choice for refining the Fe3 B/Nd2Fel4 B nanocomposite microstructure.
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 295

6.4.2 a-Fe/Nd2Fe14B System

ex-Fe/Nd2Fe14 B nanocomposites show higher coercivity than Fe3 B/Nd2Fe14 B,


thus it is worth exploring the effect of Cu and Nb additions to ex-Fe/Nd2Fe14 B
nanocomposites as well. One shortcoming of the ex-Fe/Nd2Fe14 B system is the
difficulty in controlling the microstructure. This is because the glass forming
ability of the alloy within the composition range that produces ex-Fe/Nd2Fe 14B
nanocomposites is very poor. Because of this, the production of
ex-Fe/Nd2Fe14 B nanocomposite materials in an industrial scale with good
reproducibility is challenging. There are a few investigations that reported the
beneficial effect of the addition of Cu to ex-Fe/Nd2Fe14 B nanocomposites
(Chiriac and Marinescu, 1998; Hamano et al., 1998), but no work has
clarified the role of Cu in the evolution of ex-Fe/Nd2Fe14 B nanocomposite
microstructures. The optimized hard magnetic properties of ex-Fe/Nd2Fe14 B
nanocomposites are commonly obtained by partially crystallizing the melt-spun
ribbon during solidification. The hard magnetic properties obtained after
annealing the melt-spun ribbon is very sensitive to the cooling rate of the melt-
spinning; thus the reproducibility of the hard magnetic properties obtained
from ex-Fe/Nd2Fe14 B is not as good as that of Fe3 B/Nd2Fe14 B. A good example
of the difficulty of optimizing the microstructure in the ex-Fe/Nd2 Fe14 B
nanocomposite was nicely demonstrated by Hamano et al. (1998) in the
coercivity change as a function of the wheel-surface velocities of the Cu roll for
the melt-spinning process as shown in Fig. 6. 23. The as-melt-spun ribbons
show the highest coercivity at a wheel-surface velocity of 10 m/s, while the
highest coercivity is obtained from the ribbon melt-spun at wheel surface
velocities ranging from 15 to 20 m/s after anneal ing. This is because the
sample with the highest coercivity in the melt-spun condition grows on
annealing, hence the coercivity after annealing decreases. On the other hand,
the ribbons melt-spun at faster cooling rates contain a high number density of
fine ex-Fe and Nd2Fe14 B nanocrystals embedded in the amorphous matrix;
hence, further annealing leads to an optimum ex-Fe/Nd2Fe14 B nanocomposite.
However, the microstructure of the melt-spun ex-Fe/Nd2Fe14 B nanocomposite is
very heterogeneous. Because of this, the reproducibility of the properties
obtained from melt-spun ribbons is not good. If an optimized nanocomposite
microstructure can be processed by crystallizing an amorphous precursor, the
reproducibility of the magnetic properties obtained from the ex-Fe/Nd2 Fe14 B
nanocomposites should be significantly improved. In view of this, Ping et a!.
(2000) studied the effect of Cu in the microstructural evolution process of the
ex-Fe/Nd2Fe14 B nanocomposite from the amorphous precursor.
Figure 6.24 is a 3DAP Cu map in a volume of -17 nm x 17 nm x 60 nm in
a NdsFessBsNblCUl alloy annealed at 495C for 30 min. It is apparent that Cu
atoms form clusters at this stage with a density of about 1024 m- 3 . The size of
296 Kazuhiro Hono

500 As melt-spun
~ 400
.
~ 300
:2 200
100
OL--------'------===!......- - -....

Heat-treated 180 s at 740-780 'C


500
:: 400
~ 300

:2 200
100

o 10 20 30
Ysurr.(m/s)

Figure 6. 23 Coercivity change as a function of wheel surface velocity of as-melt-spun


NdsFe765CosB6Nb,Cuo.5 ribbon and those obtained after optimal heat treatments (Hamano
et ai., 1998).

Figure 6.24 3DAP Cu map of NdsFeS5 B5Nb, CUI melt-spun amorphous alloy annealed at
495'C for 30 min.

the Cu clusters is around 3 nm. The concentration in Cu clusters was estimated


to be about 13 at. % and Nd was also sl ightly enriched in these clusters.
Although Cu atoms form clusters prior to the crystallization of this alloy, it was
found that the addition of Cu was not effective to refine the nanocomposite
microstructure. As discussed in a previous section, Cu clusters develop to fcc-
Cu to act as heterogeneous nucleation sites for a-Fe primary crystals in the
Fe-based soft magnetic materials. However, Cu clusters do not develop to the
fcc-Cu in the Nd-Fe-B alloy; hence, even if Cu form clusters, these do not
serve as heterogeneous nucleation sites for a-Fe. It is only effective to trigger
Fe3 B primary crystals. Hence, Cu addition is not useful in the a-Fe/Nd2 Fe14 B
nanocomposite.
To obtain a refined nanocomposite microstructure, the starting melt-spun
ribbon must be partially crystallized. In such a case, both Nb and Zr are known
to be effective in increasing the hard magnetic properties of the
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 297

o:-Fe/Nd2Fe 14 B nanocomposite. Figure 6. 25 shows MH corves for the


NdBFeB7 Bs , NdBFeBsBsZr2' NdBFeB4BsZr3 alloys. By the addition of 2 at. % Zr,
12

10
--- Nd sFeS7BS '"
- NdsFessBsZr2 h
....... Nd sFeS4BSZr3 0
8 ..... .'
.'.'
G'
6 6
/ ........
~ i
~
'T
4
.... f
.'
f
.' .' f
2 f
/
0 .'
-10 -8 -{j -4 -2 0
H(kOe)

Figure 6.25 Demagnetization curves of optimized a-Fe/Nd2Fe14 8 nanocomposites


produced from NdaFea7 8 s , NdaFeas 8 s Zr2' and NdaFes4 8 sZr3 alloys.

coercivity increases significantly (Wu et aI., 2001 d). By increasing the Zr


content to 3 at. %, the coercivity is substantially increased at an expense of
remanence. The microstructure change by the addition of Zr is distinct as
shown in Fig. 6. 26. The micrograph of the Zr-free alloy shows distinct
morphology difference between the Nd2Fe14 B grains and the o:-Fe intergranular

(a) (b)

Figure 6. 26 TEM bright field images of optimally heat-treated (a) Nds FeS? 8 s and (b)
NdsFess8sZr2 alloys.

grains. In the NdBFeBS BSZr2 alloy, both o:-Fe and Nd2Fe 14 B grains show
equiaxed granular shape without clear distinction. 3DAP analysis result
showed that Zr is partitioned in the Nd2Fe14 B phase with a weak tendency of
segregation at the o:-Fe/Nd2Fe14 B interface as shown in Fig. 6. 27. Although
298 Kazuhiro Hono

the partitioning tendency of Zr would be determined with EDS using a field


emission type TEM, the detection of the weak segregation of Zr at the ex-Fe/
Nd2 Fe14 B interface would not be possible with any other techniques. As seen
from this example, the 3DAP technique can provide such unique information
regarding the local chemistry of nanocrystalline alloys .
.
Nd ,
..:'.\
1).-
~J..I
. -iii- fIlIio
I
e.I
'#.1" .. ,-
.. $,:"
-. .. .'
., . ..
Zr ..
.
'.,
-' I

.'-'-1-
I

..1 .
. . ..
I
. .
. ."
..
:. , ....., .
I

'

~~~t
"
4"
,...,. I ,F ..

14nm
(a)

1.00

0.95
'i 0.90
<a
~ 0.85
.2
'0 0.80 Fe
'"
<i:
E
.9 0.15
-<
0.10

0.05
0
2 4 6 8 10 12
Depth (nm)
(b)

Figure 6. 27 3DAP elemental map and the concentration depth profiles crossing the
interfaces of o:-Fe/Nd2FeI4B phases in an optimally heat-treated NdsFess BSZr2 alloy.

In addition to Nb and Zr, many other elements such as V (Yamamoto and


Shiozawa, 1998) and Ga are known to improve the hard magnetic properties
of the ex-Fe/Nd2 Fe14 B system. However, the underlying mechanism of these
are not well understood. For example, recent work by Wu et al. (2001a) on
the effect of V in the ex-Fe/Nd2 Fe14 B nanocomposite did not find any noticeable
difference in the microstructure by the V addition. V was found to be uniformly
dissolved in the microstructure without being partitioned in either of the
phases; hence, the reason why V addition is effective in increasing coercivity
is not clear.
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 299

6.4.3 Amorphous Remaining a-Fe/Nd2Fe14B Nanocomposite


Typical nanocomposite microstructures optimized for exchange-spring magnets
do not contain any amorphous phase. However, recent work by Inoue et al.
( 1995) reported that an a-Fe/NdzFe14 B nanocomposite magnet of an
approximate composition of FeS9 Nd7B4 contains a remaining amorphous phase
in the optimal microstructure, which exhibits magnetic properties of
B f = 1.22 T, He = 240 kA/m and (BH)max = 130 kJ/m 3 . Unique features of this
nanocomposite magnet are the high iron and low boron concentrations and the
presence of the intergranular amorphous phase when the optimum magnetic
properties are obtained, although such an intergranular phase is expected to
cause a weaker exchange interaction between hard and soft magnetic grains.
Recent micromagnetic calculation results by Fukunaga et al. (1998) suggested
that a reduction of the strength of the intergrain exchange interaction between
the hard grains suppresses the averaging effect of the magnetocrystalline
anisotropy and increases coercivity and energy product. Thus, the
nanocomposite microstructure containing a remaining amorphous phase may
lead to a new composition range for Nd-Fe-B based exchange-spring magnets.
To improve the magnetic properties of the a-Fe/NdzFe14B nanocomposite
with a remaining amorphous phase, Hamano et al. (1998) studied the
microalloying effect of various additives and reported that the nanocomposite
produced from a NdsFe76 Cos NbzB6 melt-spun alloy exhibits good hard magnetic
properties of B,=1.12 T, He =512 kA/m and (BH)max=143 kJ/m 3 . A
bonded magnet prepared from this ribbon exhibited reasonably good
performance having He = 465 kA/m and (BH)max =64.1 kJ/m 3 . In addition to
Nb, they investigated the effect of Cu on the magnetic properties, but they did
not find any beneficial effect in the Cu addition. The ineffectiveness of Cu
addition to this alloy can be explained based on the conclusion in the previous
section that Cu clusters do not work as heterogeneous nucleation sites for a-Fe
as described in the previous sections.
In this alloy, however, Nb addition improves magnetic properties. Wu et
al. (2000) investigated the role of the addition of Nb in the alloys with
compositions of NdsFe76sB6CosNb, Cuos and NdsFenCosB6Nb, using the 3DAP
technique and found that the intergranular amorphous phase remains even after
optimal annealing for 3 min at 730C as shown in Fig. 6.28. Nb atoms strongly
enriched in the intergranular remaining amorphous phase as shown in the 3DAP
analysis result. It is surprising to find some of the amorphous phase remains
even after annealing above 730C . The presence of such remaining amorphous
300 Kazuhiro Hono

phase is thought to control the grain growth during melt-spinning and post-
annealing. At the same time, they would weaken the exchange force between
the hard magnetic grains, which may lead to a higher coercivity (Shrefl et aI.,
2000).

(a) (b)

,
a-Fe
, ,
Nd2Fel4B amo

0.8 Fe :
I

0.6
0.4
0.2
B ::? 0 '----,-="-T---'------J'----;>==
i 0.4 10 15
0.3
.~
Co 0.2
Iiu
0.1

~
:
.'
I
~
' .

I
", ~ U
o I

Cu ~.
~ ~.
:: . ':::::::t:=:":
15 10 15 I I

O.2~
Co: : :
0.1 I

Cu I : :
o
5 10 15
Depth (nm)

(c) (d)

Figure 6. 28 TEM bright field image and atom probe concentration depth profiles of melt-
spun NdgFe76s86CogNblCuo.s alloy annealed at 1013K for 3 min. Two types of interfaces
are analyzed with 3DAP: remaining amorphous (amo) as intergranular phase and a sharp
a-E.e/Nd, Fe!, B interface.
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 301

6. 5 Roles of Microalloyed Elements in


Nanocrystallization

As demonstrated in the previous sections, the 30AP technique is very useful


for characterizing the microstructures of nanocrystalline and nanocomposite
metallic materials. Nanocrystallization of amorphous phase often occurs by the
primary crystallization mode, in which both nucleation and diffusional growth
are involved. The crystallization of many nanocrystalline soft magnetic
materials occur by this mode. The evolution of the Fe3 B/Ndz Fe14 B
nanocomposite also occurs with the primary crystallization of Fe3 B, followed
by the polymorphous crystallization of the remaining amorphous phase.
Because of this two step crystallization process, Ndz Fe14 B crystals are
interconnected, and Fe3 B primary crystals are surrounded by the Ndz Fe14 B
grains. For making the nanocrystalline and nanocomposite microstructure less
than a 20 nm dimension, it is vital to achieve a fast nucleation rate or a large
number density of heterogeneous nucleation sites. On this basis, putting a
nucleation agent such as Cu in an Fe-based alloy is effective.
Figure 6.29 schematically illustrates the microstructure evolution
processes observed in the FINEMET type alloy and Fe3 B/Ndz Fe14 B
nanocomposites containing a small amount of Cu and a combined addition of Cu
and Nb. In the case of the FINEMET type alloy, the initial amorphous phase is
a chemically uniform amorphous solid solution. By annealing, Cu clusters
appear in the fully amorphous matrix due to the phase decomposition of Fe-rich
and Cu-rich amorphous phases. The concentration of Cu in the clusters appears
to be much lower than 100 at. % at the beginning, but it gradually increases
and the clusters eventually evolve to fcc-Cu. These Cu particles act as the
heterogeneous nucleation sites for ex-Fe primary crystals, from which Nb and B
atoms are rejected. The rejected solutes are enriched in the remaining
amorphous phase, resulting in the stabilization of the remaining amorphous
phase. This retards further growth of the ex-Fe nanocrystals. Si partitions in the
ex-Fe phase as it grows, ending up with Fe-20Si with the 00 3 structure.
Softmagnetic materials are used after the primary crystallization stage,
keeping the intergranular amorphous phase.
In the case of the Fe-TM-B (NANOPERM) type alloy, no nucleation agent
such as Cu is required. The composition of the amorphous alloy is more Fe-
rich, hence, the glass forming ability of the precursor amorphous phase may
not be as good as FINEMET type alloys. Because of this, a medium range
ordered (MRO) region with the same structure as ex-Fe develops during
solidification, which probably acts as quenched-in nuclei for the primary
crystallization of the ex-Fe phase. The nanocrystallization of this alloy also
302 Kazuhiro Hono

Hetero"eneous
Clustering of Cu nucleation of a-Fe Primary
,------.-------,
amorphous amorphous
r:-::..:..-.:...-=------, ,---,--,-----,

I
a-Fe
Cu cluster fcc-Cu
Nb & B enriched
amorphous
(a) Fe-13.5Si-9B-3Nb-l Cu(FINEMET)

Clustering of Heterogeneous
Cu & Nd nucleation of Fe3B Primary Polymorphous
,--------, ,------, ,----'---------, ,------,

+ + +
I I
Amorphous Cu & Nd cluster

(b) Fe-4.5Nd-18.5B-O.2Cu

Clustering of Heterogeneous
Cu & Nd nucleation of Fe3B Primary Eutectoid
,--------," F-------, ,--------, ('-------,

+ +
1
I
+-----J
'--+--
Amorphous Cu & Nd cluster

(c) Fe-4.5Nd-18.5B-l Nb-O.2Cu

Figure 6.29 Schematic illustrations of the nanocrystalline and nanocomposite


micorstructural evolution processes of (a) Fe-Si-B-Nb-Cu, (b) Fe-Nd-B-Cu and (c) Fe-
Nd-B-Nb-Cu alloys.

occurs by the primary crystallization of the ex-Fe phase. The remaining


amorphous phase is stabilized by the enrichment of Zr and B that are rejected
from the ex-Fe primary crystals.
In the Nds FeS6 B5Nb l and Nd s FeS5 B5Nb l CUI amorphous alloys, the Cu
atoms form a high number density of clusters (size is about 3 nm and the Cu
concentration is around 13 at. %) in the amorphous phase prior to the
crystallization of the Fe3 B primary phase. Since Nd is attractive with Cu, Nd
atoms are also enriched in the Cu clusters. By the enrichment of Nd, a local
chem ical composition close to Fe3 B is achieved near the cluster/amorphous
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 303

interfaces, from which Fe3B is nucleated (Fig. 6. 29b). Since Nd atoms are
rejected from the Fe3 B primary crystals, the Nd concentration in the
amorphous phase increases. Then, the remaining amorphous phase
crystallizes to the Ndz Fe14 B phase by the polymorphous crystallization. Thus,
the Ndz Fel4 B grains are interconnected in the microstructure. The
interconnected feature of the Ndz Fe14 B grains was clearly observed by 3DAP in
the case of the Fe3 B/Ndz Fel4 B nanocomposite as shown in Fig. 6. 30. In this
map, the Nd enriched region corresponds to the Ndz Fe14 B phase, which are
interconnected in the three dimensional space. In the case of Cu-and Nb-
containing alloy, an additional phase, FeZ3 B6 , is formed in the second stage of
the crystallization by the eutectic crystallization mode (Fig. 6. 29c). Because
of this, the grain size after the second crystallization stage is refined compared
to that of the Fe3 B/Ndz Fe14 B two phase alloy .

350
:
<>::
C.

~
~
:g
300
250
200
/ ...


Br
He!
(BH),nax
1.2

1.1

E
1.0 ~
150
~ .
sS
E
100
50
0
r
50 100 150

200 250
-
300
0.9

350
Heating rate CC/min)

Figure 6. 30 Coercivity, B" (BH )m., of Fen Ndu 8 18 . 5 Cr2 CO2 alloy as funcitions of
heating rates for annealing.

In the case of the a-Fe/Nd z Fel4 B system, the crystallization does not
progress in two stages. Even if the crystallization is started from an amorphous
precursor, the crystallization reaction occurs in one stage, suggesting that the
crystallization mode may be of eutectic type (Ping et al., 2000). Because of
this, refining the nanocomposite microstructure in less than 20 nm is very
difficult. It is also often noticed that the microstructure in the melt-spun ribbon
is very heterogeneous, containing micron scale coarse eutectic microstructures
in some parts of ribbons. In order for Cu clusters to serve as the
heterogeneous nucleation sites for a-Fe particles, Cu clusters need to develop
to fcc-Cu in the amorphous precursor as discussed in the previous section.
According to the 3DAP data, the Cu concentration in the Cu clusters observed
in the Nds FeS5 B5Nb 1 CUI amorphous alloy was only around 13 at. % , and HREM
observations did not show that there was any fcc-Cu crystallite in the
amorphous matrix. This is the reason why the Cu clusters do not work as the
heterogeneous nucleation sites for a-Fe particles during the crystallization of
the Nds FeS5 B5Nb 1CUI amorphous alloy. Thus, the formation of Cu clusters does not
have any beneficial effect on the microstructural evolution of a-Fe/Ndz Fel4 B
304 Kazuhiro Hono

nanocomposites, although it is beneficial for Fe3 B/Nd2Fe14 B nanocomposites.


It is interesting to note that the addition of Cu was reported to be effective in
improving coercivities in melt-spun Nd2 Fe14 8 single phase magnets (Herbst
et al., 1991). In this case, however, the mechanism of the coercivity
increase is completely different from what was observed in the Fe3 8/Nd2Fe 14 8
nanocomposite. Segregation of Cu along the grain boundaries was reported,
and it was believed that the grain boundary phase ping domain walls.
In the Fe3 8/Nd2Fe14 8 system, we examined the partitioning and
segregation behaviors of Co, Ga, Nb, Zr, Cu, and Cr using the 3DAP
technique. Ga and Zr partition in the Nd2Fe14 8 phase, and both tend to weakly
segregate at the Fe3 8/Nd2Fe14 8. On the other hand, Cr partitions in the Fe38
phase. Co does not show strong partitioning tendency. Nb in the
Fe3 8/Nd2Fe14 8 system causes the formation of the Fe2386 phase, to which Nb
partitions. Cu forms clusters in the early stage, which trigger the nucleation of
Fe3 8 primary crystals. Subsequently, Cu atoms partition in the Nd2 Fe14 8
phase in the final microstructure. In the ex-Fe/Nd2 Fe14 8 system, we have
examined Co, Nb, Zr, Cu, and V. Zr partitions in the Nd2 Fe14 8 phase with a
weak segregation tendency at the ex-Fe/Nd2 Fe14 8 interface. Nb is rejected
from both phases, segregating in the intergranular phase. Cu atoms form
clusters in the early stage, but these do not influence the microstructure. In the
final microstructure, Cu tends to partition in the Nd2 Fe14 8 phase. Co shows
weak partitioning tendency. Although V increases coercivity, it does not show
any partitioning tendency without having any influence on the microstructure.
As seen from these examples, understanding the roles of additives on the
microstructure and magnetic properties of nanocrystalline and nanocomposite
magnetic materials is progressing with the application of the atom probe
technique. 8y correlating the microstructure information with magnetic
properties, it has been demonstrated that better understanding of the
microalloying effect on the nanocrystalline and nanocomposite magnet are
obtained.

6.6 Effect of Heating Rates on Nanocrystalline


Microstructure Evolution

Several investigations reported that the microstructure and hard magnetic


properties of ex-Fe/Nd 2 Fe14 8 nanocomposites are sensitive to the heating rate
of the post-annealing imposed on the melt-spun ribbon. Fang and Chin (1996)
studied rapid thermal annealing (RTA) on nanocrystalline magnets with various
compositions and found that RTA is capable of yielding superior properties to
those from conventional anneal ing in the alloys with higher ex-Fe fraction. Gao
et al. (2000) reported that a higher heating rate inhibits the formation of
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 305

intermediate phases and improves the magnetic properties of melt-spun


Nd] Fe86 Nb 1B6 alloy. More recently, Kojima et al. (2000) also showed
improvement in the magnetic properties of Fe-rich nanocomposite magnets by
faster heating rate. The magnetic properties and microstructure of the
Fe3 B/Nd2Fe14 B nanocomposite magnets are also influenced by the heating rate
(Archambault and Pere, 1999). A recent report by Suzuki et al. (2000) has
also shown that the final microstructure is sensitive to the heating rate in
Nd5Fen- x Cr x B 18 (x = 0 and 3) alloys. They also showed that the addition of Cr
inhibits the formation of metastable phases such as Nd2Fe23 B3 . Figure. 6.31
shows coercivity, remanence and (BH ) max of a Fe73 Ndu B 185 Cr2 CO 2 alloy as

(a) (b)

Figure 6.31 TEM micrographs of Fe73 Ndu B I85 Cr2 CO2 alloy annealed for 10 min at 660"C
at heating rates of (a) 20 "C / sand (b) 100"C / s.

functions of heating rate (Wu et al., 2001c). When the heating rate is slower
than 100 "C/min, coercivity, remanence, and (BH)max are all lower than the
optimal value. This can be clearly explained based on the microstructural
difference as shown in Fig. 6. 32. The nanocomposite microstructure obtained

-43 nm
Figure 6. 32 3DAP Nd map obtained from the final Fe3 B/Nd2Fe" B nanocomposite
microstructure produced from a Ndu Fe75.8 B 18 .5Nb 1CU02 amorphous alloy.
306 Kazuhiro Hono

at a faster heating rate is much smaller than the slowly heated sample. They
also found that at a faster heating rate, both the Fe3 Band Nd2 Fe14 B phase
nucleate at the same time, and due to the faster nucleation rate and
impingement of the particle, the grain size can be smaller in the sample
annealed at a faster heating rate. As demonstrated here, both compositional
and processing optimization are necessary to obtain nanocomposites
microstructure.

6.7 Summary

Recent 3DAP studies on the microstructures of Fe-based nanocrystalline and


nanocomposite magnetic alloys demonstrated that this technique is extremely
powerful in providing unique information on the clustering, segregation and
partitioning behaviors of alloying elements during the nanocrystalline
microstructural evolution. Based on the knowledge of the roles of the
microalloying elements accumulated by atom probe studies, it will become
possible to design better alloy compositions more efficiently in the future. For
this purpose, it is very important to correlate the magnetic properties with the
microstructural features characterized by complimentary use of TEM and
3DAP.

References
Archambault, V. and D. Pere. Mat. Res. Soc. Symp. Proc. 577: 153
(1999)
Ayers, J. D., V. G. Harris, J. A. Sprague, W. T. Elam and H. N. Jones,
Acta mater. 46: 1861 (1998)
Bernardi, J., T. Schrefl, J. Fidler, Th. Rijks, K. de Kort, B. Archambault,
D. Pere, S. David, D. Givord, J. F. O'Sullivan, P. A.I. Smith, J. M. D.
Coey, U. Czernik and M. Gr6nefeld. J. Mag. Mag. Mater. 219: 186
(2000)
Blavette, D., B. Deconihout, A. Bostel, J. M. Sarrau, M. Bouet and A.
Menand. Rev. Sci. Instrum. 64: 2911 (1993)
Cerezo, A., T. J. Godfrey and G. D. W. Smith. Rev. Sci. Instrum. 59: 862
( 1988)
Cerezo, A., T. J. Godfrey, S. J. Sijbrandij, G. D. W. Smith and P. J.
Warren. Rev. Sci. Instrum. 69: 49 (1998)
Chen, Z., Y. Zhang, Y. Ding, G. C. Hadjipanayis, Q. Chen and B. Ma. J.
Mag. Mag. Mater. 195: 420 (1999)
Chiriac, H. and M. Marinescu. J. Appl. Phys. 83: 6628 (1998)
Coehoon, R., D. B. Demooij, J. P. W. B. Duchateau and K. H. J. Buschow. J.
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 307

de Phys. 49(C-8): 669 (1988)


Egami, T. In: Luborsky editor. Amorphous Metallic Alloys. Butterworths,
London, p. 100 (1983)
Fang, J. S. and T. S. Chin. Mater Trans JIM. 37: 883 (1996)
Fischer, R., T. Schrefl, H. Kronmuller and J. Fidler. J. Mag. Mag. Mater.
153: 35 (1996)
Fukunaga, H., T. Yamamoto, J. Kuma and Y. Kanai. Dig. 22nd Annual Conf
Magn Jpn. 96 (1998)
Gao Y, S. Zhang, B. Liu, J. Magn Magn Mater. 208: 158 (2000)
Herbst, J. F., C. D. Fuerst, R. K. Mishra, C. B. Murphy and D. J. Van
Wingerden. J. Appl. Phys. 69: 5823 (1991)
Herzer, G. Mater Sci Eng. 25: 3327 (1989)
Hamano, M., M. Yamasaki, H. Mizuguchi, H. Yamamoto, A. Inoue. In: L.
Schultz and K. H. Muller, eds. Proceedings of the 5 th International
Workshop on Rare-Earth Magnets and Their Applications.
werkstoffOlnformationsgeselischaft, Dresen, Germany, Vol. 1, pp. 199-
204 (1998)
Hirosawa, S., H. Kanekiyo, M. Uehara. J. Appl. Phys. 73: 6488 (1993)
Hirosawa, S., H. Kanekiyo, Mater. Sci. Eng. A 217/218: 367 (1996)
Hirosawa, S., H. Kanekiyo and Y. Shigemoto. Mat. Res. Soc. Symp. Proc.
577: 141 (1999)
Hirosawa, S., T. Miyoshi, T. H. Kanekiyo. Y. Shigemoto. IEEE Mag. in
press (2001)
Hirotsu, Y., M. Uehara, M. Ueno. J. Appl. Phys. 59: 3081 (1986)
Hono, K., A. Inoue and T. Sakurai. Appl. Phys. Lett. 58: 2180 (1991)
Hono, K., K. Hiraga, Q. Wang, A. Inoue and T. Sakurai. Acta metall.
mater. 40: 2137 (1992)
Hono, K., J. L. Li, Y. Ueki, A. Inoue and T. Sakurai. Appl. Surf. Sci. 67:
398 (1993)
Hono, K., D. H. Ping, M. Ohnuma and H. Onodera. Acta mater. 47: 997
(1999)
Inoue A., A. Takeuchi, A. Makino, T. Masumoto. Mater Trans JIM 36: 962
(1995)
Inoue A., Y. Miyauchi, A. Makino, T. Masumoto. Mater. Trans JIM 69: 2128
(1996)
Kanekiyo, H., M. Uehara, S. Hirosawa. Mater. Sci. Eng. A 181/182: 868
(1994 )
Kajiwara, K., K. Hono, S. Hirosawa. Mater. Trans. In press (2001)
Kojima A, A. Makino, A. Inoue. J. Appl Phys. 87: 6576 (2000)
Mak inc A, T. Hatanai, A. Inoue and T. Masumoto. Mater. Sci. Eng. A226-
228: 594 (1997)
Makino, A., T. Bitoh, A. Kojima, A. Inoue, T. Masumoto. J. Appl. Phys.
87: 7100 (2000)
Miller, M.K. Atom Probe Tomography. Kluwer Publishing (2001)
308 Kazuhiro Hono

Muller, E. W., J. A. Panitz and S. B. McLane. Rev. Sci. Instrum. 39: 83


(1968)
Nakamura, M., Y. Hirotsu, K. Anazawa, A. Makino, A. Inoue, T.
Masumoto. Mater. Sci. Eng. A 179/180: 487 (1994)
Ohkubo, T., H. Kai, D. H. Ping, K. Hono, Y. Hirotsu. Scripta mater. 44:
971 (2001)
Ohnuma, M., K. Hono, S. Linderoth, J. S. Pedersen, Y. Yoshizawa and H.
Onodera. Acta mater. 48: 4783 (2000)
Ohnuma, M., K. Hono, T. Abe, H. Onodera, S. L inderoth, J. S. Pedersen,
Y. Yoshizawa. Proc. Riso International Conference (2001)
Ping, D. H., K. Hono, S. Hirosawa. J. Appl. Phys. 83: 7769 (1998)
Ping, D.H., K. Hono, H. Kanekiyo, S. Hirosawa. J. Appl. Phys. 85: 2448
(1999a)
Ping, D. H., K. Hono, H. Kanekiyo, S. Hirosawa. Acta mater. 47: 4641
(1999b)
Ping, D. H., Y. Q. Wu, H. Kanekiyo, S. Hirosawa, K. Hono. In: H.
Kaneko, M. Homma and M. Okada eds. Proc. 16 th Inter. Workshop on
Rare-Earth Magnets and Their Applications. The Japan Institute of Metals,
505 (2000)
Ping, D. H., Y. Q. Wu, K. Hono, M. A. Willard, M. E. McHenry, D. E.
Laughlin, D. E. Scripta. Mater. submitted (2001a)
Ping, D. H., Y. Q. Wu, K. Hono. J. Mag. Mag. Mater. submitted (2001b)
Sakurai, M., M. Matsuura, S. H. Kim, Y. Yoshizawa, K. Yamauchi and K.
Suzuki. Mater Sci Eng. A 179: 469 (1994)
Schrefl, T., J. Fidler, D. Suss. In: K. Kaneko, M. Homma, M. Okada eds.
Proceedings of the Eleventh International Symposium on Magnetic
Anisotropy and Coercivity in Rare-Earth Transition Metal Alloys. Japan
Institute of Metals, p. S57 (2000)
Suzuki, K., N. Kataoka, A. Inoue, A. Makino and T. Masumoto. Mater
Trans JIM 31: 743 (1990)
Suzuki, K., A. Makino, N. Kataoka, A. Inoue, T. Masumoto. Mater. Trans.
JIM 32: 93 (1991)
Suzuki K., A. Makino, A. Inoue and T. Masumoto. Sci. Rep. RITU A 39: 133
( 1994)
Suzuki K., J. M. Cadogan, M. Uehara, S. Hirosawa, H. Kanekiyo. Scripta
mater. 42: 487 (2000)
Tate, B. J., B. S. Parmar, I. Todd, H. A. Davies, M. R. Gibbs and R. V.
Major. J. Appl Phys. 83: 6335 (1998)
Uehara, M., S. Hirosawa, H. Kanekiyo, N. Sano, T. Tomida.
NanoStructured Mater. 10: 151 (1998)
Warren, P. J., I. Todd, H. A. Davies, A. Cerezo, M. R. J. Gibbs, D.
KendaliandR.V. Major 41: 1223 (1999)
Willard, M. A., D. E. Laughlin, M. E. McHenry, D. Thoma, K. Sickafus, J.
O. Cross, V. G. Harris. J. Appl. Phys. 84: 6773 (1998)
Atom Probe Characterization of Microstructures of Nanocrystalline. . . 309

Willard, M. A., M. Q. Huang, D. E. Laughlin, M. E. McHenry, J. O. Cross,


V.G. Harris. J. Appl. Phys. 85: 4421 (1999)
Wu, Y.Q., D.H. Ping, K. Hono, M. Hamano and A. Inoue. J. Appl. Phys.
87: 8658 (2000)
Wu, Y.Q., H. Yamamoto, K. Hono. Scripta mater. In press (2001a)
Wu, Y. Q., T. Bitoh, K. Hono, A. Makino, A. Inoue. Acta Mater. In press
(2001b)
Wu, Y. Q., D. H. Ping, B. S. Murty, H. Kanekiyo, S. Hirosawa, K. Hono.
Scripta Mater. In press (2001 c)
Wu, Y. Q., D. H. Ping and K. Hono, unpublished work (2001d)
Yamamoto H. and M. Shiozawa. In: Proceedings 16 th International
Workshop on Rare-earth magnets and Their Applications. Dresden,
Germany, p. 307 (1998)
Yamasaki, M., H. Mizuguchi, M. Hamano, T. Kobayashi, K. Hono, H.
Yamamoto and A. Inoue. In: H. Kaneko, M. Homma, M. Okada eds.
Proc . 16 th Inter. Workshop on Rare-Earth Magnets and Their
Applications. The Japan Institute of Metals. pp. 513 - 520 (2000)
Yoshizawa, Y., S. Oguma and K. Yamauchi. J. Appl. Phys. 64: 6044
(1988)
Yoshizawa, Y. and K. Yamauchi. Mater. Trans. JIM 21: 307 (1990)
Yoshizawa, Y. Mater Sci Forum 307: 59 (1999)
Yoshizawa, Y. (2001) unpubl ished work
Zhang, Y., K. Hono, A. Inoue, A. Makino and T. Sakurai. Acta Metall.
Mater. 44: 1497 (1996a)
Zhang, Y., K. Hono, A. Inoue and T. Sakurai. Scripta Mater. 34: 1705
(1996b)
Zhang, Y., K. Hono, A. Inou and T. Sakurai. Appl. Phys. Lett. 69: 2128
( 1996c)
Zhang, Y., U. Czubayko N. Wanderka, V. Naundorf, F. Zhu and H.
Wollenberger. Scripta Mater. 44: 263 (2001)

This paper summarizes the work carried out in the author's group in collaboration with Dr. D.
H. Ping, Dr. Y. Q. Wu and Dr. M. Ohnuma of NIMS, Dr. Hirosawa of Sumitomo Special
Metals, Dr. Y. Yoshizawa of Hitachi Metals, Professor Inoue at IMR, Tohoku University and
Professor Makino at Akita Prefectural University. This work was supported by the Special
Coordination Fund for Promoting Science and Technology on "Nanohetero Metallic Materials"
from the Ministry of Education, Culture, Sports, Science and Technology.
7 Itinerant-Electron Metamagnetism

Kazuaki Fukamichi

7. 1 Introduction

There are many materials which exhibit the first-order transition caused by
changing external parameters such as magnetic field, pressure and
temperature as well as internal parameters such as exchange field and
composition. The first-order transition,. which takes place between a
nonmagnetic state and a ferromagnetic state, is called the itinerant-electron
metamagnetic transition (IEMT). This transition is in contrast to the transition
in localized electron magnets which are antiferromagnetic or hilimagnetic in the
grand state. This phenomenon is closely correlated to the magnetic instability
and often observed in exchange-enhanced Pauli paramagnets. This magnetic
instability causes not only the IEMT but also various striking properties such as
enhanced magnetic susceptibility and its temperature maximum, large
electronic specific heat coefficient, significant magnetovolume effects and so
on. Therefore, the IEMT is interesting from not only fundamental but also
practical viewpoints. Exchange-enhanced Pauli paramagnets, Laves-phase
compounds such as YC0 2 and LuC02 , have been investigated from both
theoretical and experimental viewpoints. In these compounds, the IEMT is
closely correlated with the peculiar band structure near the Fermi level and
spin fluctuations. Thus far, the itinerant-electron metamagnetism has mainly
been investigated from the viewpoint of basic researches. This chapter sheds
light on the marked changes in magnetic properties caused by the itinerant-
electron metamagnetic transition, intending practical appl ications.
The contents are based mainly on our recent data because extremely
comprehensive reviews on the itinerant-electron metamagnetism of lanthanide-
Co intermetallics (Duc and Goto, 1999), and formation of 3d-moments and
spin fluctuations in some rare-earth-Co compounds (Duc and Brommer, 1999)
have been performed. More recently, itinerant-electron metamagnetism and
peculiar magnetic properties observed in 3d and 5f intermetallics have been
reviewed (Goto et al., 2001).
This chapter is organized as follows. At the beginning of next Section 7.2,
theoretical researches for IEMT are briefly explained. Section 7. 3 presents
the relationship between the transition field and the susceptibility maximum for
Itinerant-Electron Metamagnetism 311

exchange-enhanced Laves-phase paramagnets. Next, the significant


magnetovolume effects in Laves-phase IEMs are discussed in Section 7. 4. The
determination of the Landau coefficients is given in Section 7. 5. In
Section 7. 6, the suppression of spin fluctuations is presented. Various
magnetic properties at finite temperatures for La (Fe x Si 1 - x) 13 are reviewed in
Section 7. 7. In Section 7. 8, drastic changes of magnetic and electrical
properties are discussed from practical viewpoints. Finally, concluding
remarks are given in Section 7.9.
The Curie temperatures in the first- and second-order transitions are
defined as T 01 and T 02' respectively, in this chapter. However, for
simplicity , To is used as the Curie temperature throughout, except for the
discussion on the magnetic phase diagram.

7. 2 Theoretical Aspects of Itinerant-Electron Metamagnetism

In this section, itinerant-electron metamagnetism is outlined. We restrict


ourselves within the discussions on essential magnetic properties. More
detailed discussions on these properties will be given in each section.

7. 2. 1 Landau Expansion Coefficients and Magnetic Phase Diagram


Landau-type expansion for the magnetic part of the free energy IlF ( M) is
given by

(7. 1)

where M is the magnetization, and a and b are the expansion coefficients. In


Eq. (7. 1), the Stoner condition for the onset of ferromagnetism is given by
a <0 for itinerant-electron systems. However, Wohlfarth and Rhodes (1962)
showed that the ferromagnetic state may be induced by applying a magnetic
field even though a> 0 if b < o. In other words, the magnetic field-induced
transition from the paramagnetic state to the ferromagnetic state takes place.
Such a transition has been called IEMT. They pointed out that IEMT occurs
when the temperature dependence of magnetic susceptibility exhibits a
maximum. Shimizu (1965) subsequently discussed IEMT by adding the M 6
term to Eq. (7. 1). In the ground state, IlF (M) of itinerant-electron systems
is expressed by

(7.2)
312 Kazuaki FUkamichi

He showed that IEMT takes place under the conditions 0>0, b<O and c>O
with 3/16 < oc / b 2 < 9/20. The coefficients 0, band c are the Landau
expansion coefficients. The value of 0 corresponds to the inverse
susceptibility at T = 0 K. Figure 7. 1 illustrates IlF ( M) as a function of
magnetization M (Shimizu, 1982). Three insets show the corresponding
magnetic curves. When oc/b 2 ~9/20, the state is paramagnetic and the
magnetization monotonically increases with the increase of applying magnetic
field. The state is metastable when 3/16<oc/b 2 <9/20, and can be
stabilized by an external magnetic field and IEMT from the paramagnetic state
to the ferromagnetic state takes place at a critical magnetic field Be,
accompanied by a discontinuous magnetization curve as shown in the inset.
The ferromagnetic state is stable when the free energy curve has two minima
at M = 0 and another one at a finite value of M under the condition of 3/ 16~
oc / b 2 , showing a representative curve of ferromagnetism. NMR experiments
have revealed that spin fluctuations play an important role in magnetic
properties of LuC0 2 and Lu (Co 1 - x Al x )2 Laves-phase compounds (Yoshimura
et aI., 1987a, 1988).

(I) (2)(3)

.......... ....
.... ,,'

Figure 7. 1 Schematic figure of the relation between the magnetic part of the free energy
AF (M) and the magnetization M. The insets show the corresponding magnetization curves
(Shimizu, 1982). (1) paramagnetic, (2) metamagnetic, (3) ferromagnetic.

Recently, theories for spin fluctuations (SFs) have been developed to


explain various magnetic properties at finite temperatures. When SF are taken
into consideration, the state can no longer be regarded as magnetically
uniform. Therefore, one must use the following free energy density per unit
volume f(r), instead of the free energy in Eq. (7.2), for an itinerant-electron
magnet (Shimizu, 1981; Yamada, 1991, 1993):
Itinerant-Electron Metamagnetism 313

fer) = ~ a I 2
mer) 1 + ~ b I mer) 14 + ~c I mer) 16 + ~ J I V mer) 12
(7.3)

where mer) is the magnetization density, a, b, c are the Landau expansion


coefficients, and J is the exchange stiffness constant. The total free energy is
obtained by integrating fer) over the whole volume V.
According to the theories for SFs, the coefficients a, band c at finite
temperatures are renormalized by thermal SFs and b.F ( M) is given by
(Moriya, 1986; Yamada, 1991)

b.F(M) = +A(nM 2 + ~ B(nM 4 + ~c(nM6, (7.4)

where ~(n2 is the mean-square amplitude of SFs, which is the sum of zero-
point and thermal fluctuations. Ignoring the contribution from the zero-point
fluctuation, ~p (n 2 is in proportion to T 2 at low temperatures and to T at high
temperatures (Moriya, 1986). The temperature dependence of a, band c is
usually very weak because the degenerate temperature in the Fermi
distribution functions is very high. Therefore, quantitative magnetic variations
with temperature can be discussed. The Landau coefficients A (n, B (n and
C ( n are functions of a, b, c, and ~p (T)2, and given as

A (n = a + ; b~ p( n 2 + 35 c~ p( n 4 ,
9
B(n = b + ~4c~p(n2, (7.5)

C(n=c.
The conditions of A (n >0, B (n <0 and C (n >0 with 3/16 <A (n:
C(n/B(T)2 <9/20 are essential for IEMT at finite temperature. On the
other hand, the ground state becomes ferromagnetic under the condition of 0<
ac/b 2<3/16. The theoretical magnetic phase diagram under the condition of
a> 0, b < 0 and c> 0 in the vicinity of the onset of ferromagnetism is
presented in Fig. 7. 2 (Moriya, 1986). The ordinate corresponds to the
temperature axis because c~p (n 2 / I b I also varies in proportion to T 2 at low
temperatures. The abscissa is the measure of the concentration x, because
ac / b 2 depends on x. The second-order magnetic phase transition (SOT)
between the paramagnetic and the ferromagnetic states occurs on the T C2 line,
and the first-order magnetic phase transition (FOT) between the paramagnetic
and the ferromagnetic state takes place on the T C1 line. The FOT disappears
on the To line. Therefore, the FOT takes place between the To and T Cl lines.
In the figure, the ferromagnetic transition changes from the second-order in the
region 0~ac/b2~5/28 to the first-order in the narrow region 5/28<ac/b 2<
3/16. Moreover, IEMT occurs just above the Curie temperature T Cl in the
narrow region, and T C1 steeply increases with decreasing ac / b 2 in the region
5/28<ac/b 2<3/16.
314 Kazuaki Fukamichi

0.3

- --
0.1
Meta.

o'--_--'-_---''--'---_-''--'--_-----'--_-------'
0.16 0.17 0.18 0.19 0.20 0.21
ac/b 2
Figure 7.2 Theoretical magnetic phase diagram under the condition of
0>0, b<O, c>O, for the Landau expansion coefficients by considering spin
fluctuations CMoriya, 1986).

7. 2. 2 Paramagnetic Susceptibility Maximum in the Temperature


Dependence
Inverse paramagnetic susceptibility, X ( T, B) - I in zero field B = 0 is equal to
A(T) given by Eq. (7.5). In the case of 0>0, b<O and c>O, X(T) is
mainly governed by the first term in Eq. (7.5) and shows a T2 -dependence at
low temperatures. The value of ~ ( T)2 exhibits a marked increase with
increasing temperature and reaches an upper limit restricted by the charge
neutral ity condition at a certain temperature T' . The system behaves as a
localized moment system above T* , and, hence the magnetic susceptibility
X (T) exhibits a Curie-Weiss type temperature dependence at high
temperatures.
The temperature T max' at which X ( T max) reaches a maximum is given by
() A ( T) 1() T. Since () A ( T) 1() T is governed by ~ ( T) 2, the temperature
dependence of 0 can be ignored, and, hence the susceptibility attains a
maximum at the temperature where () A ( T) 1() ~ ( T) 2 = 0 (Yamada, 1991,
1993), resulting in ~(T max/ = - 3b/14c and X( T max) = (a - 5b 2 128c) - I . For
oc 1b 2 <5/28, X ( T max) would become negative. Therefore, the condition for
the appearance of a maximum in the X (T) curve, X max ( T), is 0 >0, b <0,
c>O and oclb 2 <5/28. Temperature dependence of A (T) and B (T) is
schematically shown in Fig. 7.3 (Yamada, 1993). The curve of A (T) exhibits
a minimum at T max where B (T) becomes zero, that is, X (T) = A (T)-I
takes a maximum value.
Itinerant-Electron Metamagnetism 315

A(T)

Figure 7. 3 Schematic figure of the relation among the Landau coefficients


at finite temperatures, A ( T ) , B ( T) and the maximum temperature T ma> of
susceptibility (Yamada, 1993)

7.3 Itinerant-Electron Metamagnetism of Laves-Phase


Exchange-Enhanced Pauli Paramagnets

YCo2 , LuCo2 and ScCo2 Laves-phase compounds have been regarded as


representative exchange-enhanced Pauli paramagnets. Therefore, theoretical
and experimental studies for these compounds have been carried out most
extensively. In this section, the IEMT of their quasi-binary systems is
discussed.

7. 3. 1 Metamagnetic Transition in the Ground State


The IEMT is associated with a sharp peak in the density of states (DOS) just
below the Fermi level E F (Cyrot and Lavagna, 1979; Yamada et a!., 1984)
which comes from the condition b <0, that is

N' (EF)}2 _ {N"(EF) } (7.6)


{ N(E F) 3N(E F) < 0

where N' ( E F) and N" (E F) are the first and second derivatives of the Fermi
level E F with respect to DOS. This condition is satisfied when the DOS curve
is convex downward. Figure 7.4 illustrates the total DOS calculated by the
316 Kazuaki Fukamichi

linear muffin-tin orbital method for YCo z (Aoki and Yamada, 1989). Obeying
Eq. (7.6), the DOS curve at E F is convex downward. In such a band
structure, the exchange spl itting energy IlE F = IE: - E;; I is not so high as
given in the inset. Here, E: and E;; are the Fermi levels of up- and down-spin
bands, respectively. The magnitude of the DOS after the transition to the
ferromagnetic state tends to satisfy the following relation for the IEMT,
DOS(En + DOS(EF') > 2DOS(E F ) . (7.7)

100

80

T
60
1il
T
~

1
~
40

~
:<
20

O'--_ _=LL- L- -'-------L_--'

0.2 OJ 0.4 0.5 0.6


E (Ryd)

Figure 7.4 Total density of states calculated by the linear muffin-tin orbital (LMTO)
method for YC02 (Aoki and Yamada, 1989). The inset is a schematic density of state for
the itinerant-electron metamagnetic transition.

The critical transition field Be of the IEMT has been demonstrated to be


about 69 T for YCoz (Goto et al., 1989) and 74 T for LuCo z (Goto et al.,
1990). These values are close to the values calculated by using the electronic
structure of d-electrons in the magnetic field (Yamada et aI., 1987). At finite
temperatures, both compounds show a broad maximum X max in the temperature
dependence of magnetic susceptibility X ( T). By replacing Co with a non-
magnetic element such as AI, Ga, Si, and Sn, both Be and the susceptibility-
maximum temperature, T max ' are significantly decreased (Sakakibara et aI.,
1990a; Murata et al., 1991, 1993a, 1993b, 1994a, 1994b). A linear relation
between Be and T max in Y (Co 1 - x Alx)z compounds has been pointed out
(Sakakibara et al., 1990b), implying that the susceptibility maximum X ( T max)
is relevant to the shape of the DOS.
In quasi-binary systems, Y (Co l - x Alx)z (Aleksandyan et aI., 1985;
Itinerant-Electron Metamagnetism 317

Sakakibara et aI., 1990b), Lu(Co 1 - x Al x )2 (Sakakibara et al., 1987; lijima


et aI., 1990) and Lu(Co l - xGa x )2 (Saito et aI., 1997), the ferromagnetic state
becomes stable above a critical concentration x e. The magnetization curves at
4.2 K for Lu(Co 1 - xAl x )2 compounds are shown in Fig. 7.5 (Yokoyama et aI.,
2001). The data at 10K for the compounds with x = 0.020 and 0.040 are also
given by the dotted curves in the same figure, together with those of LuC0 2
(Goto et al., 1990; Goto et al., 1994). A clear IEMT with a large hysteresis is
observed. The value of Be defined as the average of the lower and higher
critical fields decreases with increasing x, keeping. almost the same
magnetization jump in magnitude. It has been pointed out that Lu(Co 1- xGa x )2
also exhibits almost the same IEMT behavior (Yokoyama et al., 2001).
1.0

0.8

/.'!
0.040
'-" /,.::.0.020 x=O
~~:::r~.~

'0 0.6 ..:',/


,
/
...
.,
:
:' :
~
;'
2,
...."

....
'

~ 0.4
.
: ::
..
..
0.2

o 20 40 60 80 100
B (T)

Figure 7. 5 Magnetization curves for Lu( Co 1- x Al x ) 2 in the concentration range from x =


0.020 to 0.090 (Yokoyama et ai., 2001), together with the curve of LuC02 for comparison
(Goto et al., 1989). The data obtained at 4.2 K are given by the solid curves, and those
at 10K by the dotted curves.

7. 3. 2 Relationship Between the Susceptibility Maximum and


the Transition Field
Figure 7.6 depicts the temperature dependence of X(T) for Lu(Col-xAlx)2in
3 T (Yokoyama et aI., 2001). All the curves show a maximum X max at the
temperature denoted as T max by the arrow, and T max decreases with increasing
x. We remark that the compound with a smaller X ( T max) has a higher Be as
seen from Figs. 7.5 and 7.6, implying that X ( T max) is correlated with Be
The relation between Beat 0 K and X ( T max) for the onset of
ferromagnetism is given by the following expression under the condition of ac /
b 2 = 3/16 (Yamada, 1993, 1997):
318 Kazuaki Fukamichi

2.0

Lu (Co l - x AIJ2
B=3T
~X=I
1.5
()()Q~0.085
~ 0.080 ....
......
~
~ 1.0
x
~

o 100 200 300 400


T(K)
Figure 7. 6 Temperature dependence of susceptibility, X ( T), for Lu ( Co , - x Al x ) 2
in a magnetic field of 3 T. The result of LuC02 is given by the dotted line (Yokoyama
et aI., 2001)

(7.8)

with

;'3TbT (7.9)
~p(Tmax) = \!~

where ~p(Tmax) and X(O) are the root mean amplitude of spin fluctuations at
T max in zero field B = 0 and the susceptibility at 0 K, respectively. Given that
~ p ( T max) is constant, Eq. (7. 8) leads to a Iinear relation between Be (0) and
X ( T max) -1. The experimental temperature 4. 2 K is low enough to regard the
observed Be as Be (0) for all the compounds.
The relation between Be and X ( T max) -1 for various Co-based Laves-
phase compounds is plotted in Fig. 7.7 (Saito et a!., 2000), together with
that for Co( S,- x Sex) 2 pyrite-type compounds (Goto et al., 1997), CeRu2 Sb
heavy-fermion compound (Mignot et al., 1988) and UCoAI wide gap-type
compounds (Mushnikov et al., 1999), for comparison. The last two
compounds are also exchange-enhanced Pauli paramagnets with X max' and
CeRu2Sb (Mignot et a!., 1988) exhibits an IEMT-like behavior, and UCoAI
(Mushnikov et a!., 1999) does a clear IEMT. It is worth noting that the B e -
X(T max )-1 plots for Co-based Laves-phase and Co(S)-X Se x )2 pyrite-type
compounds follow a universal straight solid line with the slope of about 0.4 J..Is/
Co, indicating that ~p ( T max) is almost constant irrespective of the kind of
compound systems. The observed universal Be - X (T max) -) line extrapolated
to X ( T max) -) = 0 gives a negative Be' suggesting that the condition of a >0,
Itinerant-Electron Metamagnetism 319

90

Lu (CoAI)2
Lu (CoGa)z
Lu (CoSih
"
0
Lu (CoSn)2
Y (CoAI)z
60 0 Y (CoGa)z

+ Co(SSe)z
x UCoAI
E
~o lIE CeRuzSi z

30
. E1.5~
~
On 1.0
cd
E
~0.5
1'1
::f o 0.2 0.4 0.6
q:>(,uB/mag .atom)

o 100 200 300


X(Tmaxf ' [T/(,uB/mag. atom)]

Figure 7. 7 The critical field of the metamagnetic transition, Be' obtained at 4.2 K
CB<45 T) and 10 K CB ~45 n
versus the inverse susceptibility at the susceptibility-
maximum temperature, XCTmax)-', for LuCCo 1 - x M x )2 and YCCo 1 - x Nx )2 with x = 0-
O. 09 (M = AI, Ga , Sn and Si, N = AI and Ga) (Saito et al., 2000), together with that for
YC02(Gotoet aI., 1989), LuC02 (Goto et aI., 1990), Y(Col-xAlx)2(Be~45 n
(Sakakibara et al., 1990a), Co( SI- xSe x )2 with x = O. 12-0.20 (Goto et aI., 1997b),
CeRu2 Sb (Mignot et aI., 1988) and UcoAI (Mushnikov et aI., 1999). The inset shows that
the magnetic moment induced by applying magnetic field, Mind' versus the slope of the
B c - X(T max )-1 relation, <p.

b<O and c > 0 holds just above the critical ferromagnetic composition
because this condition results in X ( T max) (Yamada, 1993). As is well known,
ScC02 is also an exchange-enhanced paramagnet and shows X ( T max) in a
similar manner as YC0 2 and LuC02 However, no IEMT has been confirmed in
magnetic fields up to 120 T (Sakakibara et aI., 1990b). Assuming the same
universal Be - X ( T max) -1 relation, Be of ScC02 is estimated from the data of
X (T) (Ishiyama et aI., 1984) to be about 150 T, nearly equal to the
theoretically calculated value from the electronic structure of d-electrons in the
external magnetic field (Terao and Yamada, 1987). 0 ifferent linear Be -
X ( T max) - 1 relations for CeRU2 Sb (Mignot et al., 1988) and UCoAI (Mushnikov
et aI., 1999) are also confirmed as given by the dotted lines. The slope of the
Be - X ( T max) - 1 relations, cp, for the former, and the latter compounds are
320 Kazuaki Fukamichi

about O. 49 J.JB/Ce and O. 14 J.JB/U, respectively, different from the above-


mentioned value for Co-based Laves-phase and Co (SI-X Se x )2 pyrite-type
compounds.
The value of the magnetization induced by applying a magnetic field at
OK, M ioo (0), can be approximately expressed as [3 I b I I (4 c) ] 1/2 in the
vicinity of ac I b 2 = 3/16, and Eq. (7.8) is rewritten as

(7.10)

Therefore, quantitative discussion is possible by using Eq. (7. 10) because


Mind is obtainable from the magnetization curve. The resultant values of Mind at

4. 2 K for Co-based Laves-phase and Co (SI- x Sex) 2 pyrite-type compounds


are 0.6 ...... 0.8 J.JB/Co, in agreement with the value of 0.7 J.JB/Co estimated
from Eq. (7. 10). This impl ies that X ( T max) is strongly related to b.-F (M) in
the ground state. It is noteworthy that Eq. (7. 10) is valid for the compounds
with a high Be such as YCo 2 and LuCo 2 , as shown in Fig. 7. 7, though the
equation is applicable in the vicinity of aclb 2 = 3116 which is the condition of
the onset of ferromagnetism. The inset in Fig. 7. 7 shows the Mind vs. rp plot
for Co-based Laves-phase and Co (SI- x Sex) 2 pyrite-type compounds (Goto
et aI., 1997), CeRu2 Si 2 (Mignot et aI., 1988) and Ueo AI (Mushnikov et al.,
1999) . A linear relation between Mind and rp is observed. The slope of the
straight line Mindl (Bel X ( T max) -1) is about 2, comparable with 40/21 given in
Eq. (7.10), implying that X(T max ) of CeRu2Si2 heavy-fermion compound and
UCoAI wide gap-type compound is also closely connected with b.-F (M) in the
ground state.

7. 3. 3 Metamagnetic Transition at Finite Temperatures


The critical field Be of the metamagnetic transition at finite temperatures is
given by the following expression (Yamada, 1993):

B
e
=-.!Jlbl{a-l~}+-.lJlbllbl~(n2.
3 3c 16 c 16 3c p
(7.11)
Figure 7.8 shows Be vs T 2 for Lu(Co - xAl x )2(Fukamichi et al., 2001). These
'
compounds exhibit an excellent linear relationship. Therefore, following Eq.
(7. 11), Be is directly proportional to T 2 in the low-temperature region. The
value of Be for the compounds with x = O. 020 and O. 040 in Fig. 7. 5 is
respectively estimated to be about 70 T and 51 T at 4.2 K, about only O. 2 T
lower than at 10K. Consequently, the data at 10K are almost the same as
those at 4. 2 K.
Itinerant-Electron Metamagnetism 321

12

10

E 6
c:C;

o 2 3 4 5 6
T 2 (X 103K 2)

Figure 7. 8 The critical transition field Be against the square of temperature r 2 for
Lu( Co 1- x Al x ) 2 in the concentration range from x = O. 085 to O. 098 (Fukamichi et al.,
2001)

7. 4 Correlation Between the Magnetovolume Effects


and Metamagnetic Transition

7. 4. 1 Concentration Dependence of the Curie Temperature and


Spontaneous Magnetization
Shown in Figs. 7.9 and 7. 10 are the concentration dependence of the Curie
temperature T c and the spontaneous magnetization M s for Lu(Co 1- x Al x )2 and
Lu(Co 1- x Ga x )2 systems, respectively (Saito et aI., 1997, 1999; Yokoyama
et al., 2001). Two systems are almost the same in the concentration
dependence and the ferromagnetic state is suddenly developed at Xc ( Saito
et aI., 1997, 1999; Yokoyama et aI., 2001). Strictly speaking, the values of
Lu(Co 1- x Al x )2 are slightly higher than those of Lu(Co 1- x Ga x )2. Furthermore,
it is interesting to note that the concentration with the highest T c does not
coincide with the concentration where M s exhibits the largest value as seen
from Figs. 7.9 and 7. 10. In other words, the increase in T c and the decrease
in M s are observed around x = O. 100 -- O. 150 with increasing x. As is well
known, these behaviors are characteristic to conventional Fe-Ni Invar alloys
and Invar-type Fe-Pt and Fe-Pd alloys (Wassermann, 1990), as well as
322 Kazuaki Fukamichi

0.8 , . - - - - - - - - - - - - - - - - - , 150

0.6 -=-
100
E
~
(3 0.4
Co
-3
~'" 50
0.2

o'--_---&-..L-_ _-'---_ _--'-_ _---' 0


0.05 0.10 0.15 0.20 0.25
x

Figure 7.9 Concentration dependence of the spontaneous magnetization Ms at 4 2 K and


the Curie temperature T c for Lu(Co , - x Al x )2 (Yokoyama et aI., 2001).

0.8 150

0.6 -=-
100
E
~
(3 0.4
0,
g
-3 ~
'i 50
0.2
,.I Lu (Col_xGaxh

I
a a
0.05 0.10 0.15 0.20 0.25
x
Figure 7. 10 Concentration dependence of the spontaneous magnetization Ms at 4. 2 K
and the Curie temperature T c for Lu (Co, - x Ga x ) 2 (Saito et al., 1999; Yokoyama et al.,
2001) .

amorphous Invar alloys CFukamichi, 1983; Fukamichi et al., 1989).

7. 4. 2 Pressure Effects on the Curie Temperature and


Spontaneous Magnetization
Figure 7. 11 shows the concentration dependence of the pressure derivative of
Te,JTc!JP, for LuCCo'-xAlx)2andLuCCol-xGax)2CSaito et aI., 1999;
Itinerant-Electron Metamagnetism 323

Yokoyama et aI., 2001). The magnitude of a T cia P for both compound


systems is very sensitive to the concentration, and a significantly large
negative value is observed in the concentration toward the onset of
ferromagnetism. The concentration dependence of aT 0/ a P for both compound
systems is comparable to that for conventional and amorphous Invar alloys
( Wassermann, 1990; Fukamichi, 1983, 1989). We should recall, in
connection with Figs. 7. 9 and 7. 10, the similarity of the concentration
dependence of M s and To among Lu (Co'-x Al x )2' Lu (Co 1- x Ga x )2'
conventional and amorphous Invar alloys.
120

100
Lu (Col_xMxh

- - M=AI
-0- Ga

40

20

O'-----L---L---'----'----'-------''------'
0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22
x
Figure 7. 11 Concentration dependence of the pressure derivative of the Curie
a~ .
temperature,JP for Lu(Co l - x Al x )2 and Lu(Co l - x Ga x )2(Salto et al.. 1999; Yokoyama
et al.. 2001).

To explain the pressure dependence of To, the magnetovolume effects as


well as spin fluctuations should be taken into account. By considering the
magnetovolume coupling energy, the Landau coefficients should be replaced
and the magnetic part of the free energy t.F ( M) of itinerant-electron systems
is written as (Yamada, 1995)

t.F(m) = -.loM 2+ -.l6M 4 + -.lcM 6 (7.12)


246
c
with 0 = a + 2KC mv P, 6 = b - 2KC~v and = c, where the coefficients 0, 6
and c are functions of a, b, c, and K and C mv are the compressibility and the
magnetovolume coupl ing constant, respectively. The conditions of 0> 0, 6 <
o and c> 0 with 3/16<oc/6 2 <9/20 are necessary for the IEMT. A negative
b is related to a positive curvature of the DOS at E F as well as the negative
mode-mode couplings among SFs. At finite temperatures, the coefficients in
Eq. (7. 12) are renormal ized by thermal SF. Recently, it has been pointed out
324 Kazuaki Fukamichi

that the region of the FaT at T Cl becomes wider by considering the


magnetovolume coupl ing energy. That is, the FaT takes place under the
following condition (Yamada et a!., 2002):
5 ac 3
28 - TJ < 52 < 16 (7.13)

with

_ 2 2
TJ - --_-KC rnv (7.14)
71bl
Figure 7. 12 shows the magnetic phase diagrams with TJ = O. 01 and 0
which are given by the solid and dashed lines, respectively, corresponding to
the states with and without the magnetovolume effect (MVE) (Yamada et a!.,
2002). As seen from the figure, the magnetovolume coupling energy widens
the region of the FaT at T Cl. The type of the ferromagnetic transition changes
to the SOT under the condition of O<ac / 6 2 <5/28 - TJ. In the spin fluctuation
theory, the pressure dependence of T C under the conditions of 0, 6 < 0 a>
and c>0 can be discussed by considering the pressure effect on the mean-
square amplitude of SFs at TC2,~p(Tc2/. Therefore, a~p(TC2)2/ap is given
by the following expression (Yamada et al., 2002):

a~p(TC2)2 =_ 3 KC rnv (~_ TJ _ ac2 + 2TJ2)-1/2 (7.15)


aP J35 I 6 I 28 b 5
The value of a ~p ( T C2)2 /a P is proportional to a n2 /(j P, because ~p ( T C2)2 is
0.3
T
' .........c ,\
Para.

0.2 k: ..

Ferro. Meta.

0'---------'------"-----'-----
0.16 0.18 0.20
aell?
Figure 7. 12Theoretical magnetic phase diagram under the condition of < 0, fj < 0, a
C>0, for the Landau expansion coefficients by considering spin fluctuations and
magnetovolume effects (Yamada et aI., 2002).
Itinerant-Electron Metamagnetism 325

directly proportional to n2 at low temperatures ( Yamada, 1993). Equation


(7. 15) implies that a remarkably large negative value of J T ed3 P is observed
in the vicinity of the critical concentration Xc for the onset of ferromagnetism
because the condition of oc / iJ2 = 5/28 - 1) is close to 3/16 at xc. In a similar
way,
-- 1

J~p(Tel)2 =_ ~ k~mv [~C _ ~ + 1) + 281)2J-2 (7.16)


3P -!7 I bib 2 28

The theoretical values of 3 ~ p ( T e2 ) 2/3 P and 3 ~ p ( Tel) 2/3 P as a function of


oc Iij2 are given in Fig. 7. 13(Yamada et aI., 2002). A prominent dependence
is accompanied by a divergent-type cusp at 5/28 - 1). The experimental data
on 3 Te/d P given in Fig. 7. 11 are qualitatively consistent with the theoretical
results. Note that the increase of x corresponds to the decrease of oc / 6 2,
that is, Fig. 7. 11 is equivalent to the view of the reverse side of Fig. 7. 13.
Consequently, the observed large negative value of J T e/ J P given in Fig. 7. 11
can be explained by the conditions of 0>0,6<0 and c>O with oc/6 2 =
5/28 - 1).

3/2Cmv

o 5/28-71 3/16
ac/fi 2
Figure 7. 13 Pressure dependence of the first-order transition temperature
T Cl and the Curie temperature T C2 as a function of the Landau coefficient ratioae/ b2
(Yamada et aI., 2002).

Figure 7. 14 shows the concentration dependence of the pressure


coefficients of 31nT e /dP and JlnMj3P for Lu(Co 1 - x Al x )2 and Lu(Co 1 - X Ga x )2
(Saito et aI., 1998, 1999; Yokoyama et aI., 1999, 2001). Both of the
coefficients for the former system are essentially the same as those of the
latter system. The magnitude of 3lnM s /d P is proportional to Ke mv as well as
X hI and given by the following equation (Wagner and Wohlfarth, 1981).
326 Kazuaki Fukamichi

2.5

Lu (Co _x Mx12
'
2.0
- - InM~/P
.,
~
M=AI
- - InTel?
0
'"
~

\.5
~ Ga
--0- InMsI?
~ -0- InTelP
.5
"" 1.0
~

~ 0.5

0
0.08 0.12 0.16 0.20 0.24
x
Figure 7. 14 Concentration dependence of the pressure coefficients, 31nTcl3 P and
3InM s /3P, for Lu(Co - x Al x )2 and Lu(Co - x Ga x )2(Saito et al., 1998, 1999; Yokoyama
' '
et aI., 1999, 2001).

(7.17)

where Xhf is the high-field susceptibility. The value of Ke rnv is estimated to be


about 4--7X 1O- 3 ps-2. It should be noted that the magnitude of alnTc/ap is
much larger than a InMs/a P below around x = O. 150, showing a rapid
decrease of T c compared with that of Ms. Those results indicate that M s
begins to decrease drastically above a certain pressure, closely related to the
MT induced by applying pressure. Therefore, the sign of 6 for the compounds
with x ~O. 150 is negative, implying that the positive curvature of the DOS at
E F remains up to around x = O. 150 (Fig. 7. 4). In other words, these results
imply that there is a sharp peak of the DOS just below E F up to around x =
O. 150.
Concentration dependence of the ratio of aln Tc;a P to alnMs/a P, that is,
alnTc/alnM s for Lu(Co'-xAU2 and Lu(Co'-xGax)2 systems is plotted against
the concentration x in Fig. 7. 15 (Saito et al., 1999; Yokoyama et al., 2001) .
The value of a InTc;a InM s for both compound systems increases with
decreasing x and becomes about 7 for Lu (Co'-x Al x )2 with x = O. 100.
According to the theory for weakly ferromagnetic itinerant-electron systems
with the conditions of 0 < 0, b > 0 and c = 0, aIn T cia InM s is unity for the
Stoner-Wohlfarth theory (Wohlfarth, 1969) and 3/2 for spin fluctuation theory
(Takahashi, 1990). The results under the condition 0>0, b<O and c>O, on
the other hand, in the vicinity of the onset of ferromagnetism are much larger
than 3/2. It should be emphasized that a InTc/a InM s is given by only the
Landau coefficients of a, c
6 and (Yamada, 1993; Yamada et ai., 2001).
Itinerant-Electron Metamagnetism 327

The magnitude of a InTc/a InM s becomes larger with increasing 2


oc/b
, or

decreasing x close to the onset of ferromagnetism. A more detailed discussion


of this behavior will be given in Section 7.7.3.
10

- - M=AI
-0- Ga

0'---------"----------'----------'---------'
0.08 0.12 0.16 0.20 0.24
x
Figure 7. 15 Concentration dependence of a InTcia InM s for Lu (Co , - x Al x )2 and
LU(Co'-xGa x )2 (Saitoetal., 1999; Yokoyama et aI., 2001).

7. 4. 3 Thermal Expansion Anomaly and Spin Fluctuations


The magnetic contribution to the thermal expansion is given by

(7.18)

where Ws ( T) is the spontaneous volume magnetostriction at T. The


temperature dependence of each term is presented in Fig. 7. 16. The curve
associated with the term of ~p (T)2 is convex downward at low temperatures
and exhibits a linear increase at high temperatures, whereas M ( T) 2
decreases and disappears at T c with increasing temperature as shown by the
dot-dashed line. The value given in the figure for M(0)2 in P = 0 is obtained
from the following equation:

M(0)2 = I b_ I (1 + 4_0C).
2
(7.19)
2c b
The schematic thermal expansion curve is given by the solid line in the same
figure.
The value of t::.ws/ Ws (0), defined by the following expression, is plotted
against the Landau coefficients oc/b
2
in Fig. 7.17 (Yamada and Terao,
1994):
328 Kazuaki Fukamichi

"-
Sp (Tef ---------------~\-----~
... ,
\ .... i
.. \ :
Sp(T)2 . \ i
)....... \ :
........ :
Tc T

Figure 7. 16 Schematic explanation of the relationship among the thermal expansion


anomaly (= spontaneous volume magnetostriction) Ws (T), mean-square amplitude of
spin fluctuations ~p (T)2 and magnetization M( T).

1.00

S
~ 0.75
N

~
'V'
I
N 0.50
Moriya_1
.5 II
I
I
I
S 1
I

~ 0.25 I
1
I
S I
I
<l I
I
I
I
a
5/28-1] 3/16
ac/[J2
Figure 7.17 Relative change of the spontaneous volume magnetostriction !:>.ws/!:>.ws (0)
against the Landau coefficient ratio 2
ac/5
(Yamada and Terao, 1994). The value obtained
by the Moriya theory (Moriya, 1985) is indicated by the arrow, for comparison.

!::J.Ws = [MCO)2 - ~pC T C )2J


(7.20)
wsCO) MCO)2

where wsCO) and MCO) are the values at 0 K. At oc/b 2 =5/28, !::J.ws/wsCO)
in the case of positive mode-mode couplings, i. e., 0<0, b>O of Eq. (7. 1)
is O. 4 CMoriya, 1985), whereas the value is O. 72 in the case of negative
mode-mode coupl ings, i. e., 0> 0 and b < 0 in Eq. C7. 2). Furthermore.
Itinerant-Electron Metamagnetism 329

LlWs/WsCO) goes up to 1.0 with reaching ac/b 2 to 3/16. This means that a
marked reduction in the thermal expansion coefficient occurs in the vicinity of
the FOT and SOT magnetic phase transitions given in Fig. 7.12.
Thermal expansion curves of Lu CCo 1 - x Ga x )2 with x = O. 09, O. 10 and
O. 11 are presented in Fig. 7. 18 CHayashi et al., 2000). It should be noted that
a marked thermal expansion anomaly is confirmed by changing the
composition, namely, the curve of the paramagnetic compound with x = O. 09
exhibits no anomalous thermal expansion, whereas the curves ferromagnetic
compounds with x = O. 10 and O. 11 show a significant anomalous temperature
dependence below around T c, denoted by the arrow. Such anomalous thermal
expansion characteristics are explained by using Figs. 7. 16 and 7. 17. That is
to say, the magnetization M ( T) decreases and disappears at the Curie
temperature, while the thermal SFs are excited, and hence p CT)2 increases
with increasing temperature. Therefore, the smaller the thermal expansion
coefficient, the stronger the temperature dependence of M CT) and/or the
weaker the increase of thermal SFs.
a
o
a
t:" x=O.11 a
0.10 a
a
o 0.09 a
o
a
C1
a
Tc a Q

~. "f .ao
....r ~o
.........~ 0

o 0
0000

o 50 100 150 200 250


T(K)
Figure 7. 18 Thermal expansion curves of Lu( Co 1- x Ga x )2 with x = 0.09, O. 10 and O. 11
(Hayashi et al., 2000).

7. 5 Determination of the Landau Coefficients

7.5. 1 Pressure Effect on the Critical Field of the Metamagnetic


Transition

As discussed in the preceding Sections 7. 4. 2 and 7. 4. 3, the large


magnetovolume effects are observed in Lu CCo 1 - x Al x ) 2 and Lu CCo 1 - x Ga x ) 2 .
330 Kazuaki Fukamichi

The magnetovolume effects influence not only the pressure dependences of T c


and Ms but also Be of the MT. Figure 7. 19 shows the pressure dependence of
Be at 4.2 K for LUCC00900Alol00)2 and LuCCo0900 GaolO0)2CSaito et al., 1999;
Yokoyama et ai., 2001). The value of Be increases linearly with the pressure
as seen from the figure, and the value of a Be/J P is estimated to be
9. 2 T/ GPa for Lu CCoo 900 Al o 100 ) 2' sl ightly smaller than 12 T/ GPa for Lu
CCo0900GaO 100)2' The critical pressure PI defined as the pressure where the
transition field becomes zero is estimated to be 0.43 GPa for Lu( Coo. 900 Al o.100)2
and 0.11 GPa for LuCCoO.900 Gao.100)2 by a linear extrapolation to Be=O.
12

M=Oa AI

E
ci{

4.2K

0 0.8 1.2 1.6


P (OPa)

Figure 7. 19 Pressure dependence of the critical transition field Be at 4. 2 K and the


calculated value of Beat 0 K for Lu (Coo 900 Al o 100) 2 and Lu (Coo. 900 Gao. 100 ) 2 (Saito et ai.,
1999; Yokoyama et ai., 2001).

The effect of pressure on the width of hysteresis t:.B e defined as the


difference between the lower and higher critical fields for Lu (Coo 900 Al o 100) 2
and LuCCoogOOGaOI00)2 is given in Fig. 7.20 (Saito et al., 1999; Yokoyama
et al., 2001). The critical pressure P 2 at which the first-order MT disappears
is estimated to be 2. 1 for the former and 1. 3 GPa for the latter by a linear
extrapolation. As seen in Figs. 7. 19 and 7.20, both PI 0.43 and P 2 2. 1 = =
GPa for Lu(CoogooAI0100)2 are larger than the values of PI =0. 11 and P 2 = 1. 3
GPa for Lu(CoogooGao 100)2' These results imply that the ferromagnetic state of
the former is more stable than that of the latter, in accordance with the higher
T c in Fig. 7.9, compared with the data in Fig. 7.10.
The magnetoelastic coupl ing constant, KC mv' and the Landau expansion
coefficients for Lu CCoo. 900 Al o.100 ) 2 and for Lu CCoo 900 Gao. 100 ) 2 are estimated
from the experimental results given in Figs. 7. 19 and 7. 20. The effect of
thermal SFs on the free energy is negligible because the magnetization
=
measurement temperature T 4.2 K is low enough to ignore the thermal SFs.
Itinerant-Electron Metamagnetism 331

1.5

,, 4.2 K
,
1.0

E
cil
<l

0.5 ,,
,,
,, P2
,, P2

\~ " ~
0 0.5 1.0 1.5 2.0
P(GPa)

Figure 7.20 Effect of pressure on the width of hysteresis of the critical


field Be at 4.2 K for Lu (COO.900 AI 0. loo )2 and Lu (COO 900 GaO.100 )2 (Saito et aI., 1999;
Yokoyama et aI., 2001).

By considering the magnetovolume effect, the following term should be added


to the equation of state in the magnetic system (Yamada, 1995):
l!. F mv =- Gmv wM ( T) 2 (7 . 21>
where w is the relative change of volume, T is the temperature,

w = V ~oVo =- KP + KG mv [M(T)2 + ~p(T)2J (7.22)

where Vo is the volume when M = 0, K is the compressibility. The magnetic


equation of state B ( w) is written as
B(w) = o.(P)M + 6M + eM 3 5
. (7.23)

The Landau expansion coefficients a and b are modified by the


magnetovolume coupling as given in Eq. (7. 12). It should be noted that only
the coefficient a in Eq. ( 7. 12) is affected by the pressure through the
pressure dependence of the volume. The value of Be becomes zero when the
value of o.e/6 2 = 3/16, and the FOT disappears when o.e/6 2 = 9/20
(Yamada, 1993; Shimizu, 1982). Therefore, the observed values of both PI
and P 2 are connected with the following equations, respectively:
ac
-2 - (a + 2 KG mv PI) c 3 (7.24 )
6 (b - 2KG~v)2 -16
and
ac
-2 - (a + 2 KG mv P2 ) c 9 (7.25)
6 (b - 2KG~v)2 - 20'
332 Kazuaki Fukamichi

From Eq. (7.23), the pressure effect on M~ at T = 0 is given by the following


equation

Mt = I 6_ I [1 +
2c
J
1_ 4 ac ].
62
(7.26)

The measured M~ of the compound with x = 0.10 is 0.46 (J.1s/CO)2 (= 6.0 X


10 4 emu2/cm 6 ) at T = 4.2 K under ambient pressure. From Eqs. (7.24),
(7.25) and (7.26), the values of KCmv/O(=o), 6/0(=J.1) and c/o(=\fJ)
and can be determined. Using the experimental result of Be under pressure,
the coefficient 0 is obtainable because two minima of /;;.F (M) given by the
following equation are equal each other at Be:

/;;.F(M) = ~ (0 + 2ooP)M 2 + ~ J.10M 4 + ~ \fJoM 6 - MB e . (7.27)

Then, the values of 6, C and KC mv are calculated by using the estimated value
of 0 and the available value of the compressibility K = 8.5 X 10- 13 dyn/cm 2 for
LuCo 2(Klimker et aI., 1979). The estimated values of the Landau coefficients
a ,6 ,c and KC mv for x = O. 100 are Iisted in Table 7. 1, together with those of
LuCo2(Yamada et al., 1995; Goto et aI., 1998), Co(SogSeOI)2(Goto et aI.,
1997) and Fe3 Pt (Entel and Schroter, 1989; Sumiyama et aI., 1981), for
comparison. The value of /;;.F ( M) for all the compounds and Fe3 Pt has two
local minima in the paramagnetic and the ferromagnetic states. The energy in
the ferromagnetic state is lower than that in the paramagnetic state, indicating
that the ferromagnetic state is stable in the grand state. One may expect a
first-order ferromagnetic transition at T c from those results. According to the
Landau expansion as mentioned above, the FOT at P = 0 occurs under the condition

Table 7. 1 Estimated values of the Landau coefficients b, a, c


and the magneto-elastic
coupling KC mv ,!), 5/28-1) for LU(COO.90oAloIOO)2 and LU(Co090oGaolOO)2(Yokoyama et aI.,
2001), together with data for LuC02(Yamada, et aI., 1985; Goto and Bartashevich, 1998),
Co( Se09 Seo 1) 2(Goto et aI., 1997) and Fe3 Pt (Sumiyama et aI., 1981; Entel and Schrbter,
1989) .
abc
-- -2 KC mv
(10'cm'l (lO-'cm'l (lO-'cm' 1 aclb (lO-'~B-') I) 5128-1) Reference
emu) erg) erg')
Lu( Coo '00 Al o 100)' 6.9 -7.8 1.2 0 14 66 0.011 O. 17 (1)

Lu( Coo 900 Gao 100)' 7.6 -8.3 1.4 O. 16 7.4 0.014 O. 16 (1)
LuCo, 269 -242 65 15.2 (2) ,(3)

Co(So ,Seo 1)' 2.2 -5.4 2.3 (4)


Fe,PI 4.1 -5.9 '.3 2.4 (5), (6)
(1) (Yokoyama el aI., 2001);
(2) (Yamada, 1995);
(3) (Golo and Barlashevich, 1998);
(4) <Golo el aI., 1997);
(5) (Enlei and Schroler, 1989);
(6) (Sumiyama el aI., 1981).
Itinerant-Electron Metamagnetism 333

given in Eq. (7.13). In the case of 0~ac/j)2~5/28-17' the type of the


ferromagnetic transition changes to the second-order. As seen from Table 7. 1
(Yokoyama et al., 2001; Fukamichi et aI., 2001), the value of ac/
j)2 for
LU(COO,900GaO,100)2 is estimated to be O. 16, very close to 5/28-17 . On the
other hand, the value of ac/j)2 for Lu(Coog oo Al o,lOo)2 is estimated to be O. 14,
slightly smaller than 5/28-17. The estimated value of Ke rnv for
Lu(Coo.900AI0100)2 in the paramagnetic state is very close to 6.9 X 10- 3 (IJ.B/
Co)-2 of a paramagnetic Lu(Co o 920 Gao oBo)2(Saito et aI., 1999).

7. 5. 2 Comparison Between the Experimental and Theoretical


Magnetic Phase Diagrams
Figure 7. 21 displays the thermomagnetization curves in a magnetic field of 0.5 T for
the ferromagnetic Lu( Co 1 - x Al x ) 2 with x = 0.095, 0.098 and O. 100 (Fukamichi
et al., 2001). The compounds with x = O. 098 and O. 100 are ferromagnetic
and T c is determined to be 55 and 64 K, respectively. The
thermomagnetization curves exhibit no hysteresis, indicating the SOT. On the
other hand, the curve of Lu (Coo 905 Al 0095 ) 2 exhibits a clear hysteresis around
=
the Curie temperature T Cl 28 K, and the magnetization in the vicinity of T C
decreases more rapidly than that with x = O. 098. Therefore, it is concluded
that the FOT occurs in Lu (COO,905 Al 0095 )2' In order to elucidate such
behaviors, the magnetization curves up to 3 T for Lu (COO.905 A1 0095 )2 are
presented every 2 K just above T C1 in Fig. 7. 22 (Fukamichi et aI., 2001). The
FOT becomes broader with increasing temperature, resulting in the increase of
Be and finally disappears in the temperature range between 40 and 42 K, and
hence the critical temperature To for Lu (COO 905 A1 0095 )2 is estimated to be
42 K. Such magnetic curves just above T Cl are also observed in the
paramagnetic Lu(Co l - x Al x )2 with x=0.085 and 0.090 and their To is slightly
lower than that for x = O. 095. For the onset of the metamagnetic transition,
the fourth-order term of the Landau expansion should be negative as given in
connection with Eqs. (7.2) and (7. 4). The Arrott plots just below the Curie
temperature T C for Lu ( Co 1- x Al x ) 2 compounds are presented in Fig. 7. 23
(Fukamichi et al., 2001). A negative slope is confirmed in the plots for x =
O. 095, verifying that the fourth-order term is negative.
As shown in Eq. (7. 11) and Fig. 7. 8, the observed T 2 dependence of
Be originates from the thermal SF ~p ( T)2 proportional to T 2 at low
temperatures. The slope hardly depends on x, implying that the thermal SF in
the ferromagnet where x = O. 095 above T Cl is equal to that in the
paramagnetic compounds. The value of Be for the ferromagnet where x =
O. 095 should become zero at T Cl' From this relation, T CI is estimated to be
26 K, practically in agreement with T CI obtained from the thermomagnetization
curves given in Fig. 7.21. In order to determine the critical concentration Xc
of the onset of ferromagnetism for Lu (Co 1 - x Al x ) 2 and Lu (Co 1 - x Ga x ) 2' the
334 Kazuaki Fukamichi

0.7

0.6 Lu (Co 1_xAl x)2


B=0.5 T
0.5

0
0.4
\"lell
.3 0.3
~
t\
0.2
x=0.095
0.1

0 50 100 150
T(K)
Figure 7. 21 Heating and cool ing thermomagnetization curves of Lu ( Co 1- x Al x ) 2 where
x = O. 095, 0.098 and O. 100 in a magnetic field of 0.5 T (Fukamichi et ai., 2001).

0.7
Lu (COO 905 Al 0 095)2
0.6

0.5

0 0.4
\"lell ---.-32K
.3 --0-34 K
~ 0.3
--36K
--D--38K
0.2 --+--40 K
-<>-42K
0.1 -.-44 K

0
1.0 1.5 2.0 2.5 3.0 3.5
B(T)

Figure 7.22 Magnetization curves of Lu( COO 905 A1 0095 )2 just above the Curie temperature
T Cl in magnetic fields up to 3 T (Fukamichi et ai., 2001)

concentration dependence of Be at 0 K is plotted in Fig. 7. 24 CFukamichi


et ai., 2001). As shown in the figure, Be linearly decreases with increasing
x. The value of X e at which Be becomes zero corresponds to the onset of
ferromagnetism. From this relation, the critical concentration x e for
LuCCo 1 - x Al x )2 and Lu CCo 1 - x Ga x )2 is estimated to be 0.092 and 0.096,
respectively, indicating that x = O. 095 for Lu CCo 1- x Al x )2 and O. 100 for
Lu( Co 1 - x Ga x ) 2 are very close to the critical concentration for the onset of
ferromagnetism.
Itinerant-Electron Metamagnetism 335

0.5

x=0.095
0.4

N
0.3
0'
u
ill
::l
~
0.2
~
0.1
Lu (Co l - x Al x )2

0 2 4 6 8 10 12 14
B/M(T/IlB)

Figure 7. 23 Arrot! plots for Lu(Co 1- x Al x )2. At 35 K for x = 0.095, at 62 K for x =


O. 098 and at 70 K for x = O. '00 (Fukamichi et aI., 2001).

40

Lu (Co 1_x Mx)2


30

E 20
0:;

10
- - Al
- 0 - Ga

-10
0.05 0.06 0.07 0.08 0.09 0.10
x
Figure 7. 24 Concentration dependence of the critical transition field Be at 0 K for
Lu(Co - x Al x )2 and Lu(Co 1- x Ga x )2(Fukamichi et aI., 2001).
'

Using the observed values of the Curie temperatures, T 01 and T 02' and
the critical temperature To, the magnetic phase diagrams in the x-T plane in
the vicinity of the critical concentration between metamagnetism and
ferromagnetism for LuCCo - x Al x )2 and LuCCo 1- x Ga x )2 systems are established
'
as shown in Figs. 7.25 and 7.26, respectively CFukamichi et aI., 2001). The
value of T 01 steeply increases near the onset of ferromagnetism with increasing
x, whereas To is not so sensitive to the concentration of x. The experimental
336 Kazuaki Fukamichi

phase diagrams in the vicinity of the onset of ferromagnetism for both systems
reproduce the theoretical phase diagram as seen from Fig. 7. 12. Note again
that the increase of x in both figures corresponds to the decrease of b2. ac/
Alternatively, Figs. 7.25 and 7.26 are equivalent to the view of the reverse
side of Fig. 7. 12. The ordinate in the theoretical phase diagram corresponds
to the magnitude of the mean-square amplitude of ~p (T C )2 in Fig. 7. 12.
Therefore, the rapid increase in the Curie temperature of Lu(Co 1- x M x )2 with
increasing x indicates a marked increase of ~ p ( T c) 2. Moreover, the onset of
ferromagnetism in the ground state and the metamagnetic transition above T Cl
by applying magnetic fields are also explained by the theoretical phase
diagram based on the conditions of a>
0, b < 0 and c>
O. What has to be
noted is that the FaT occurs in very narrow concentration ranges. In Table 7. 1,
the value of ac/
b 2, estimated to be O. 14 for Lu ( Coo 900 Al o 100 ) 2l is sl ightly
smaller than 5/28-1). This means that the ferromagnetic transition of this
compound is of the second-order. The experimental phase diagram given in
Fig. 7. 25 is consistent with the value mentioned above. Furthermore, the
value of ac /b 2 for Lu (Co0900 Gao 100) 2 is estimated to be O. 16, very close to
5/28 -1) under the condition between the FaT and the SOT. Note that, in
accordance with the results, the value of T Cl for Lu (COO.900 GaO. 100 ) almost
meets the To Iine as seen from Fig. 7. 26, namely, the obtained value is in
agreement with the experimental phase diagram given in Fig. 7. 26.
Therefore, it is concluded that the phase diagrams in the vicinity of the onset
of ferromagnetism for Lu(Co 1 - x AU2 and Lu(Co 1 - x Ga x )2 are explained well by
the theoretical phase diagram proposed by Yamada et al. (2001) .
100

80

60

g
h 40 To

20
I
I
0'-------'--''-------'-----'----------'
0.08 0.09 0.10 0.1\ 0.12
x

Figure 7. 25 Magnetic phase diagram of Lu (Co 1- x Al x ) 2 system in the vicinity of the


critical concentration (Fukamichi et al., 2001).
Itinerant-Electron Metamagnetism 337

100
Lu (Col-xGaxh

80

Para.
60

g
h
40
To

20
Meta.
I Tel
I
0
0.08 0.09 0.10 0.11 0.12
x

Figure 7. 26 Magnetic phase diagram of Lu (Co - x Ga x )2 system in the vicinity of the


'
critical concentration (Fukamichi et ai., 2001).

7.6 Suppression of Spin Fluctuations in Laves-Phase


Metamagnets

Itinerant-electron models and localized moment models are unified in the spin
fluctuation theory (Moriya, 1985), and various magnetic and electrical
properties over wide temperature ranges covering the paramagnetic
temperature ranges have been discussed extensively. SFs are sensitively
affected by temperature, pressure magnetic field, etc. The effects of
temperature and pressure have been discussed in the preceding Sections
7.2.2 and 7. 4. 2. In this section, the effects of the magnetic field on
suppression of SFs are discussed.

7. 6. 1 Concentration Dependence of the Specific Heat Coefficient


of Laves-Phase Compounds
SFs interact with electrons and bring about the following enhancement of the
electronic specific heat coefficient (Ikeda et al., 1991):

+ (T:J (T:J J+ t'T


2

CiT = Yth = Yband[:; 0'0 In


2
(7.28)
with
338 Kazuaki Fukamichi

(7.29)

where Ylh is the theoretical electronic specific heat coefficient, Yband is the
band term, m' / m = 1 + Aep + ASF is the mass enhancement given by electron-
phonon interactions (Aep) and spin fluctuations(AsF)' T SF is the spin fluctuation
temperature, and S T is the Stoner enhancement factor. The last term is the
lattice specific heat term and the coefficient f3 is correlated with the Debye
temperature. The experimental values of the electronic specific heat
coefficient Yexp of LuCo z (Saito and Fukamichi, 2001) and YCo z (Muraoka
et ai., 1979) are Iisted in Table 7. 2, together with the band terms Yband
calculated by the tight-binding approximation (Yamada et ai., 1984, 1985).
The value of Yexpfor both compounds is much higher than Yband determined from
the DOS, i. e., the band term, implying that the enhancement of the electronic
specific heat coefficient due to SFs is considerably large. Concentration
dependence of the Yexp values for Lu( Co 1 - xGa x ) z and Lu (Co,- x Al x ) z systems
is given in Fig. 7.27 (Saito et al., 2001; Yokoyama et ai., 2001). With
increasing x, Yexp decreases at first and increases up to the concentration in
the vicinity of the onset of ferromagnetism, and then drastically decreases in
the ferromagnetic range due to suppression of SFs, and finally increase
gradually because of a magnetic weakness around x = O. 20 (F igs. 7. 9 and
7. 10). It should be noted that the values in the ferromagnetic state are much
larger than the theoretical value given in Table 7. 2 for LuCo z ' implying that
the contribution from SFs is still large.

Table 7. 2 Experimental values of the electronic specific heat coefficient Yexp for YC02
(Muraoka et ai., 1979) and LuC02 (Saito and Fukamichi, 2001), together with theoretical
values Yband calculated by the light-binding (TB) approximation (Yamada et ai., 1984,
1985).
Compound Yexp (mJ/ (mol' K 2
12.7(2)
13.7(41
( 1) (Saito and Fukamichi. 2001);
(2) (Yamada et al.. 1985);
(3) (Muraoka et al.. 1979);
(4) (Yamada et al.. 1984).

7. 6. 2 Large Electronic Specific Heat Coefficient Due to Spin


Fluctuations and Its Suppression Under High Fields
Field dependence of the electronic specific heat coefficient Yexp for
Lu(Coo 90 Gao 09)Z is presented in Fig. 7.28, together with its magnetization
curves at 4.2 K (Saito and Fukamichi, 2001). This compound exhibits an
IEMT in 5 T, and a pronounced decrease of Yexp occurs around the
Itinerant-Electron Metamagnetism 339

50

45 --Ga
-o-AI

;=:: 40
N~

i 35
~
I 30
~
~

25

20

15 L-_----'_ _----'-_ _---'--_ _--'----_---'


o 0.05 0.10 0.15 0.20 0.25
x

Figure 7. 27 Concentration dependence of the electronic specific heat coefficient Yexp of


Lu( CO - x AI,)2 (Yokoyama and Fukamichi, 2001) and Lu (Co , -, Ga, )2 (Saito and
'
Fukamichi, 2001)

corresponding transition field. Eventually the value above 8 T is reduced to


about one-half its initial value at zero field. According to the band calculations
by the tight-binding approximation for YC0 2 in both paramagnetic and
ferromagnetic states (Yamada et al., 1984), the change in DOS at E F before
and after the metamagnetic transition is relatively small. Furthermore, the
50 , - - - - - - - - - - - - - - - , 0.8

0.6

0.2

20F='------.J----'------'------'---------' 0
o 2 4 6 8 10
B(T)

Figure 7. 28 Magnetic field dependence of the electronic specific heat coefficient Yexp and
the magnetization curves for LU(C0091 GaO 09 )2 (Yokoyama and Fukamichi, 2001).
340 Kazuaki Fukamichi

reduced value above 8 T is comparable to the value in the ferromagnetic state


in the zero field (Fig. 7.27). Therefore, the large decrease of Yexp is mainly
attributed to the suppression of SF in the ferromagnetic state induced by the
IEMT.

7. 7 Metamagnetic Transition at Finite Temperatures


of Ferromagnetic La(Fel-xSix)13 NaZn13-type
Compounds

La (Fel- x Six> 13 has a cubic MaZnwtype structure with Fm3c( O~) space-group
symmetry, and the conventional cell is composed of 112 atoms. There are two
kinds of Fe atoms: Fe! and Fen. This system becomes ferromagnetic in 0.11~
x ~O. 19 (Palstra et aI., 1983). The local environment of the Fe! atom is
very similar to that of Fe-based alloys. As a result, La(Fe l - x Six )13 exhibits
significant magnetovolume effects in analogy with Fe-Ni Invar alloys and Fe-Pt
Invar-type alloys (Wassermann, 1990).

7. 7. 1 Magnetization and Magnetic Phase Diagram


Figure 7.29 shows the concentration dependence of the Curie temperature T c
and the spontaneous magnetic moment per Fe atom M s of La (Fel- x Six) 13
(Palstra et aI., 1983; Fujita and Fukamichi, 2001b). The value of Tc
decreases, whereas M s increases with decreasing x. Such behavior is
characteristic to Invar-type alloy systems, associated with magnetic
instability. Laves-phase compound systems also exhibit a similar tendency in
the concentration of AI and Ga between about O. 10 and O. 15 as seen from
Figs. 7. 9 and 7. 10. Figure 7. 30 depicts the temperature dependence of
magnetization in 0.3 T for La(Fel-xSix)13 with x = O. 12, 0.14 and O. 16
(Fujita et aI., 2002b). In the range of low temperatures, the magnetization
decreases gradually with temperature. Although only an SOT at T c was
confirmed without any drastic anomalies in the reported previous data (Palstra
et al., 1984), a discontinuous change takes place at 195 K for x = O. 12,
indicating an FOT. Such an FOT in La(Fel-xSix)13 is limited to the higher Fe
concentration region. On the other hand, M s gradually decreases against T as
the SOT for the compounds with x~0.16. For La(Feo88SioI2)13 in Fig. 7.31,
the M6ssbauer spectrum at 193 K shows one sextet component of the magnetic
ordered state, and a paramagnetic doublet coexists with a ferromagnetic
sextet at 195 K (Fuj ita et al., 2002a). The sextet component suddenly
disappears and only the doublet component is observed at 197 K.
Consequently, the transition of the compound where x = O. 12 is not from the
Itinerant-Electron Metamagnetism 341

ferromagnetic to the antiferromagnetic state, which is often observed in Fe-


based compounds (Nishihara and Yamaguchi, 1984; Irisawa et aI., 2000), but
to the paramagnetic state. The coexistence of the ferromagnetic and the
paramagnetic states is due to a supercooling phenomenon, being the
characteristic feature of the FOT.
2.2 , - - - - - - - - - - - - - - - - - , 300

,-., 2.0 250


E
0
OJ
<l) g
~ ~
3
;;: 1.8 200

e'"
oD.
Fujita and Fukamichi
Palstra et al.

1.6 150
0.10 0.12 0.14 0.16 0.18 0.20
x
Figure 7. 29 Concentration dependence of the Curie temperature T c and the spontaneous
magnetization M sfor La(Fel-x Six) 13 (Palstra et aI., 1983; Fujita and Fukamichi, 2001 b) .

2.4

x=0.12
2.0

0.14
E 1.6
0
'i<J
<l)
'"'- 1.2
Co
3
~ 0.8

0.4

0 50 100 150 300


T(K)

Figure 7. 30 Thermomagnetization curves in O. 3 T for La (Fel- x Six) 13 (Fujita et al.,


2002a).

It is considered that the free energy as a function of magnetization M for


the compound where x = O. 12 has two minima at M = 0 and M # 0, which
respectively correspond to the paramagnetic and the ferromagnetic states
(Fig. 7. 1). Under these conditions, the energy structure gives not only the
342 Kazuaki Fukamichi

197 K
,<v~ ... II

195 K
'i2~, -;': ~'pWi
.~ ":.~/'\. i ,';\V~1
~ .. ! 2%
:~j
193 K j,' ,

x=O.12
'I.

\ to, l'~ . :'\. A


'r.'," ;,;":.,f V "1

;1~
2%

-8 -4 0 4 8
Velocity (mm/s)

Figure 7,31 Mbssbauer spectra measured at 193, 195 and 197 K for La(Feo.88 Sio 12)13

(Fuj ita et ai, 2002a).

thermal-induced FOT between the' ferromagnetic and the paramagnetic states,


but also the magnetic-field induced IEMT from the paramagnetic state to the
ferromagnetic state above T c . Figure 7. 32 illustrates the isothermal
magnetization curves for x = O. 12 around T c in order to elucidate the onset of
the IEMT (Fujita et al., 1999). The magnetization curves show a non-linear S-
shape process with an apparent hysteresis, indicating an FOT. The onset of
the IEMT is conclusive evidence in support of the double minimum structure of
free energy. The thermal energy makes the minimum in the ferromagnetic
state shallower, and the minimum in the paramagnetic state becomes stable
above T c. The magnetic field gives the Zeeman energy to the ferromagnetic
state and the IEMT occurs. Since the entropy increases more steeply in the
paramagnetic state than in the ferromagnetic one, the energy difference
between the ferromagnetic and the paramagnetic state becomes larger with
increasing temperature, resulting in the increase of Be as shown in Fig. 7.32.
In other words, the thermal induced FOT at T c corresponds to the IEMT at
Be = O.
The IEMT at finite temperatures has been discussed in terms of both the
Landau-type free energy expansion and the renormalization effect of SF on the
expansion coefficient (Yamada, 1993). For the onset of the IEMT, the fourth-
order coefficient of the Landau expansion coefficient should be negative,
giving a negative slope of the Arrott plots. In order to discuss the condition for
the IEMT, the Arrott plots for x = O. 12 at 210 K and for x = O. 14 at 220 K
are illustrated in Fig. 7.33 (Fujita et al., 1999). The negative slope is
Itinerant-Electron Metamagnetism 343

2.0,-----------------
190 K
200 K
210 K
220 K
1.5

230 K
E
0
~ 1.0
t
2>'" I 240 K
~
250 K

0.5
x=0.12
Tc=195 K

o 2 4 6 8 10 12
B (T)

Figure 7.32 Isothermal magnetization curves in the temperature range between 190 and
250 K for La(Feo.88 Si o 12) 13 (Fujita et aI., 1999)

3
x=0.12
210 K

x=0.14
220 K

o 2 4 6 8
B/M(T/PB)

Figure 7. 33 Arrot! plots for x = O. 12 at 210 K and for x = O. 14 at 220 K for


La(Fel-xSix)13(Fujita et aI., 1999)

observed in the plot for x =


O. 12 and the plot for x =
O. 14 shows no
negative slope, but has an inflection point which is also explained by the
negative fourth-order Landau expansion coefficient (Fig. 7. 23). It is worth
nothing that the origin of the negative fourth-order Landau coefficient of Laves-
phase compounds is attributed to unique Co 3d band structures (Cyrot and
Lavagna, 1979; Yamada and Schimizu, 1985). Furthermore, the IEMT may
344 Kazuaki Fukamichi

be influenced by not only the electronic structure but also the elastic energy
change (Entel and Schr6ter, 1989; Mohn and Schwarz, 1992; Yamada,
1993), because such a large increase in the magnetic moment should be
accompanied by a remarkable volume expansion. Therefore, as discussed in
connection with Eq. (7. 12) and Fig. 7. 12, the influence of the
magnetovolume coupling on the free energy should be relatively large.
Recently, the theory based on a fixed moment method by considering both the
influence of renormalization effect of SF on the free energy and the
magnetovolume effect clearly shows minima of two magnetic states in the free
energy for Fe3 Pt Invar-type alloy, namely, a ferromagnetic state with a large
volume and a paramagnetic state with a small volume (Entel and Schr6ter,
1989). This theoretical result has been mainly discussed in terms of the
symmetry and the volume dependence of 3d-electron bands rather than
pecul iar band structure (Mohn and Schwarz, 1992). A large number of Fe
atoms in La(Fel-xSix)13form icosahedral clusters which have a local symmetry
similar to that in an fcc structure (Palstra et al., 1984; Fukamichi et al.,
1994). Therefore, it is considered that the symmetry of 3d-electron bands of
La(Fel-x Si x )13 resembles that of y-Fe and fcc-Fe alloys. Recent band
calculations for La(Fel-xSix)13 (Fujita et aI., 2002c) show that the density of
states at E F is similar to that of Fe-Pt Invar alloys which are nearly strong
ferromagnets, i. e., the up-spin band is almost fulfilled in electron
occupations. Experimental data such as Xhf at 4.2 K for La(Fel-xSix)13(Fujita
et al., 2002c) are consistent with the band calculations, namely, the value of
X hI is close to that of an Fen Pt28 Invar-type alloy, in contrast to that of Fe-Ni
Invar alloys which are weak ferromagnets having holes in up- and down-spin
bands. Further experimental and theoretical investigations are necessary for
detailed discussion on the relation between the magnetic state in y-Fe and fcc-
Fe alloys and the IEMT.
The magnetic phase diagram of the La(Fel- xSix) 13 system is constructed
in Fig. 7. 34 (Fujita et aI., 2002b). The Curie temperatures T Cl and T C2
increase, whereas To decreases with increasing x, indicating that the IEMT
occurs in a wide temperature range. In theoretical calculations for itinerant-
electron systems with the negative mode-mode coupling among SF (Moriya,
1986; Yamada, 1993), the phase boundaries are derived from the critical
conditions as a function of ~p ( T)2. By assuming a proportional relation
between ~ p ( T) 2 and T, and by neglecting the concentration dependence of
the dispersion of SF, the critical point To decreases with decreasing T c. The
calculated phase diagram reproduces the experimental results given in
Figs. 7.25 and 7. 26 for Laves-phase compounds. In contrast with these
results, the data on To of La(Fel-XSix)13 increases with decreasing T c . These
different behaviors may be concerned with the concentration dependence of
the damping and dispersion coefficients of SFs, which should be measured by
microscopic dynamical experiments such as neutron scattering and NMR.
The IEMT is often followed by significant magnetovolume effects due to an
Itinerant-Electron Metamagnetism 345

260

240 0

TC2

g 220
h

TCI
200

La(Fe l_xSi x)13

180
0.10 0.12 0.14 0.16 0.18
x
Figure 7.34 Magnetic phase diagram of La( Fe,- x Six) 13 system. The Curie temperatures
of the first-order and the second-order transitions are denoted by T 01 and T C2 ,
respectively. The temperature To stands for the critical point of the metamagnetic transition
(Fujita et ai., 2002b).

abrupt change in the local magnetic moment (Figs. 7.16 and 7. 18). Figure 7.35
shows X-ray diffraction profiles for x = O. 12 measured in the temperature
range crossing the Curie temperature T c = 195 K (Fujita et aI., 2002a). No
difference between the diffraction patterns of the profiles at 140 K and 200 K is
confirmed for 6 sets of the plane indexed in the figure, although a significant
shift of the peak positions is observed. Therefore, the onset of the
ferromagnetic ordering has no influence on the symmetry of the atomic lattice,
but results in a large volume change. In more detail, the volume change occurs
discontinuously. That is, a large volume phase coexists with a small volume
phase as seen from the profiles at 170 - 190 K. The discontinuous change of
the volume and the coexistence of the large and small phases are attributed to
the FOT between the ferromagnetic and the paramagnetic phases and the
supercooling effect around the transition temperature, in accord with the
M6ssbauer spectra in Fig. 7. 31. In other words, the onset of the magnetic
moment at the transition temperature brings about a striking volume
expansion, being about 1.2 % in the temperatures of 170 - 250 K or 0.7 T c -
1. 1 T c. Such features are explained by I1F (M) described by the Landau-type
expansion in terms of the series of M as given in Eq. (7.12). The negative 6
gives the negative mode-mode coupling among spin fluctuations and has
important influences on not only the IEMT but also on the magnetovolume
effects characterized by a marked spontaneous volume magnetostriction. For
example, the first principles calculation predicts a double minimum structure of
346 Kazuaki Fukamichi

free energy as a function of M and V for Fe3 Pt ordered and disordered fcc
alloys, and the magnetovolume effects in Fe3 Pt are discussed in terms of the
energy barrier between these local minima (Entel and Schroter, 1989). It
should be noted that the theoretical calculation by a fixed spin moment method
gives an FOT in Fe3 Pt ordered alloy (Entel and Schroter, 1989), though a
martensitic transformation prevents us from observing such a magnetic phase
transition. On the other hand, La(Fel- x Six) 13 has no martensitic
transformation, and consequently yields the FOT without any changes in the
crystal structure.

220 K

210 K

200 K

195 K

190 K

180K

98 100 102 104


2en
Figure 7.35 X-ray diffraction profiles at various temperatures for La(Fe088 Si o 12) 13 (Fujita
et al., 2002a).

7. 7. 2 Thermal Expansion Anomaly

Since the IEMT above T c in La (Fe l - x Six) 13 is concerned with the double
minima structure of the free energy, as explained in connection with
Fig. 7. 32, it is expected that the transition at T c even in the zero field is also
Itinerant-Electron Metamagnetism 347

an FOT from the ferromagnetic minimum state to the paramagnetic state.


Shown in Fig. 7. 36 are thermal expansion curves for the compounds where
x=0.12, 0.14, and O. 16 CFujita et al., 2002b). In cooperation with the
change in the feature of the thermomagnetization curves in Fig. 7. 30, the
volume change around T c becomes continuous with decreasing x. The volume
decreases from low temperatures up to T c in these three specimens,
indicating the existence of the spontaneous volume magnetostriction in the
ferromagnetic state. Furthermore, the magnitude of the negative thermal
expansion coefficient seems to become smaller with decreasing x. The change
of the magnitude of the thermal expansion coefficient corresponds to the
temperature dependence of Ws C T), which is generally expressed by
Eq. (7.18). Accordingly, the significant spontaneous volume magnetostriction
in the ferromagnetic state for LaCFel-xSix)13is also attributed to the negative
mode-mode coupling among SF explained by Eq. (7.20) .

.. ~------

o 100 200 300


T(K)

Figure 7.36 Thermal expansion curves of LaCFe'-xSix)13 with x=O.12, 0.14 and O. 16
CFujita et aI., 2002b). The solid circles are the data from X-ray diffraction.

7. 7. 3 Pressure Effect on the Metamagnetic Transition


Since a pronounced volume change is caused by the magnetic phase transition
in La CFel- x Six) 13' a significant influence of the hydrostatic pressure on the
phase transition is expected. Figure 7. 37a, b show the temperature
dependence of M under hydrostatic pressure P for LaCFel- x Six) 13 where x =
348 Kazuaki Fukamichi

O. 12 and 0 14. respectively (Fujita et al . 2002b). A strong decrease of T c is


observed with increasing P. Especially. the discontinuous change of
magnetization for x = O. 14 becomes more significant with increasing P.
indicating a sharper metamagnetic transition. The pressure dependence of M
and T c is shown in Fig. 7. 38a, b, respectively (Fuj ita et al., 2002a). Both M
and T c exhibit a negative pressure dependence and the pressure coefficient (J
InM/J P is- 0.015 and- 0.019 GPa- 1 , while JlnTc/J P becomes - 0.48 and
- O. 31 GPa -1 for x = O. 12 and 0, 14. respectively. By neglecting the
influence of zero-point SFs, the pressure dependence of M at low
temperatures is mainly concerned with a change of the width of the 3d electron
band. whereas that of T c is connected with the valance between the
magnetovolume effect caused by both P and SFs (Yamada. 1995). By using
the Landau expansion coefficients, the critical pressure Pc (T) for the FOT
from the ferromagnetic to the paramagnetic state is expressed as follows
1.8

x=0.14

--
1.2 B=0.3 T
c;;
"
u.. o GPa
dl 0.27 GPa
.5
"- 0.50 GPa
~ 0.6
0.77 GPa
1.00 GPa

I I
a 50 100 150 200 250
T(K)
(a)
2.4

2.0

1.6 x=0.J2
8 B=0.3 T
'"
1.)
"- 1.2
~ --OGPa
::, 0.27 GPa
'"
<. 0.8 0.50 GPa
0.77 GPa
0.4 1.00 GPa

I
0 50 100 150 200 250
T(K)
(b)

Figure 7.37 Temperature dependence of the magnetization M measured as a function of


applied hydrostatic pressure for Ca) LaCFeo86 Sic 14 ),3 and Cb) LaCFeo88 Sio 12)13 CFujita
et al., 2002b),
Itinerant-Electron Metamagnetism 349

2.1
x=0.12
~

'" 2.0
~
.3 l--o-_~
~ 0.14 ~--~--

1.9

o O.~ ~~ ~~ l~ 1~
P (GPa)
(a)

80'-----_---'-_ _.1...-_--'-_ _-'--_----'


o 0.25 0.50 0.75 1.00 1.25
P (GPa)
(b)

Figure 7. 38 Pressure dependence of (a) the spontaneous magnetization M sand (b) the
Curie temperature T c for La(Fe'-xSi x )'3 with x=O. 12 and O. 14 (Fujita et aI., 2002a).

(Yamada, 1993):
(7.30)
where f~ and fg are the constants consisting of KC mv and the coefficients a, b
and C in Eq. (7. 12). The temperature dependence of ~ p ( T) 2 is given by

~p( T)2 = L (k B ~)2 (7.31)


o 61TQ

The parameters y and 0 are concerned with the damping and the dispersion
coefficients of spectrum of SFs (Yamada, 1993). The critical pressure at the
constant temperature in Eq. (7. 30) gives the critical temperature at the
constant pressure. As shown by the sol id lines in Fig. 7. 38b, the data are well
reproduced by Eq. (7. 30). It should be noted that the critical pressure Pc ( T)
in Eq. (7.30) is not a function of M (T) but only ~p ( T)2. In itinerant-electron
systems, it is well-known that the curve of the free energy f:!.F ( M) itself is
changed by SFs as explained by Eq. (7. 4). The hydrostatic pressure
enhances the thermal change in the curve of f:!.F (M) due to SFs, and hence a
local minimum in the ferromagnetic state becomes shallow, resulting in the
350 Kazuaki FUkamichi

lower first-order T Cl. Consequently, the influence of pressure on T C is larger


than that on M (Figs. 7.14 and 7.15).
The Landau expansion gives the following another important relation
between T c and the discontinuous change of M at T c' IlM ( T c) (Aharony,
1983; Yamada, 1993):

2 3 Ib + 14 c~ p ( T c) 21
3
IlM(T c ) = 4 c . (7.32)

Figure 7. 39 reveals the relation between Il M ( T c) 2 and n


for La (Fe 1 - x Six) 13
with x = 0.12 and O. 14 (Fujita et al., 2002a). In both the compounds, the
linear relation between IlM ( T c )2 and n
is confirmed. Accordingly, these
results reveal that the pressure effect on T c is explained well by the Landau
expansion renormalized by SFs. Validity of Eq. (7.32) strongly implies that
the pressure effect on T c is dominated by SFs.
2.5

2.0

j 1.5

~
~ 1.0

0.5

O'------'-------'---...J....:>o-----'
123 4 5
TZ(X J0 4 K2)

Figure 7. 39 Square of the discontinuous change of the magnetization !:lM at the Curie
temperature T c plotted against n obtained in various hydrostatic pressures for
La(Fe'-xSi x ),3 with x=O. 12 and O. 14 (Fujita et aI., 2002a).

7. 7. 4 Control of the Metamagnetic Transition by Hydrogen


Absorption
In contrast to application of hydrostatic pressure discussed in Section 7.7. 3,
the lattice expansion would be caused by hydrogen absorption as if negative
pressure. The room temperature lattice constant a I against the Curie
temperature T c for La (Fe088 Si o 12) 13 Hy is presented in Fig. 7.40 (Fujieda
et al., 2001). The value of a I of La ( Fel- x Six) 13 is increased by hydrogen
absorption without the change in the NaZn,rtype structure. The value of T c is
increased up to 336 K, accompanied by the volume expansion of about 1.2% .
Itinerant-Electron Metamagnetism 351

This volume expansion is caused by not only the hydrogen absorption but also
the magnetovolume effect, because La (Feo88 Sio 12) 13 H16 is ferromagnetic at
room temperature. The value of Q, in the ferromagnetic range is apparently
larger than that in the paramagnetic range. Therefore, a very large
magnetovolume effect is preserved after hydrogen absorption. For the
compound where y = 1. 0, a thermally induced phase transition occurs around
room temperature because T c = 278 K. The diffraction patterns in both the
ferromagnetic and the paramagnetic states exhibit the NaZn13 -type structure.
The peak positions show a significant shift around T c, and the coexistence of
peaks in the ferromagnetic state with a large volume and the paramagnetic
state with a small volume is confirmed at 280 K just as Fig. 7.35 shows. Such
a discontinuous peak shift suggests that a thermally induced FOT also takes
place after hydrogen absorption.
1.160

1.157
/ Ferro.

1.154
E
-=-<J 1.151

1.148

1.145 l . . -_ _- - ' -_ _- - ' - - ' - -_ _--'

150 200 300 350

Figure 7.40 The room temperature lattice constant Q, against the Curie temperature T c
for LaCFeo.ssSio.12)13 HyCFujieda et aI., 2001).

The temperature dependence of the relative volume change determined from X-


ray diffraction patterns for La(Feo.88 Sio 12) 13 HlO is illustrated in Fig. 7.41. Due to
the FOT, the volume V is changed discontinuously at T c in analogy with the
thermal expansion curve of La ( Feo. 88 Si O. 12 ) 13 (Fig. 7. 36). For itinerant-
electron magnets, the spontaneous volume magnetostriction Ws ( T) is
expressed by using the magnetization M (T) of the local magnetic moment
and ~(T)2 as given in Eq. (7.18). In the ferromagnetic state, the contribution
of thermal SFs to the volume is smaller than that of large local moments. On
the other hand, the magnetization of the local moment becomes zero in the
paramagnetic state where the volume is dominated by thermal SFs. The
difference in volume l:i.ws ( T c) is expressed by the following expression
(Yamada and Terao, 1994):
352 Kazuaki Fukamichi

La(Feo 88 Si O 12)I3 H I 0
Tc=278 K

Ferro.

1
0 . 2% Para.

250 260 270 280 290 300 310 320


T(K)
Figure 7.41 Temperature dependence of the relative volume change for
La(Feo.aaSiOI2)13 H IO (Fujieda et aI., 2001)

1.8

, ,0 ' . 0" 0 o ,0 266 K


0"0
................... 280 K
1.5 _'-.'''. o{'S-_.-6 284 K
........ -~' """-288K
~~292K
1.2 296 K

E
0
1il 300 K
0.9
'"
t.l..
dl
2>
~
0.6

0.3 Tc= 278 K

La( Feo88 Si o 12)13 H 1,0

o 2 3 4 5 6 7
B (T)

Figure 7.42 Isothermal magnetization curves at various temperatures for


LaCFeoaaSio,12)13 HIO CFujieda et aI., 2001).
Itinerant-Electron Metamagnetism 353

(7.33)

where M CT c)2 is the squared magnetization at the Curie temperature and


~p CT c )2 is the mean-square amplitude of spin fluctuations at T c . For
LaCFeos8 Sio 12) 13 H IO , the difference in volume between the ferromagnetic and
paramagnetic states is about 1%. This result implies that M CT c) becomes
smaller and ~p CT c )2 becomes larger with increasing T c. After hydrogen
absorption, therefore, the value of b.w s CT c) La CFeOS8 Sio 12) 13 H lo is smaller
than that of LaCFe088Si012)13'
Given in Fig. 7. 42 are the magnetization curves as a function of
temperature around T c for La CFe088 Sio 12) 13 H IO CFujieda et a!., 2001). A
characteristic S-shape curve with a clear hysteresis is observed above T c ,
indicating that a field-induced FOT also takes place after hydrogen absorption.
In other words, the IEMT at room temperature is real ized by hydrogen
absorption, implying that the change of the DOS curve around E F is negligible
after hydrogen absorption. From the practical viewpoint, therefore, the
hydrogen absorption is one of the useful methods to control T c .

7. 8 Drastic Changes in Magnetic and Electrical


Properties and Their Practical Applications

The IEMT brings about marked changes of various magnetic and electrical
properties. In this section, as typical examples, isotropic giant
magnetostriction, large magnetocaloric effect and giant magnetoresistance will
be discussed from the standpoint of practical applications.

7.8.1 Isotropic Giant Volume Magnetostriction


A giant magnetostrictive compound, Terfenol based on TbFe2, was
commercialized late in the 1980s, and many applications have been reported.
Recently, newtypes of magnetostrictive materials are drawing attention. One
is a large magnetic-field-induced strain due to the rearrangement of twin
boundaries of variants in the martensitic transformation for ferromagnetic
shape-memory alloys CUllakko et a!., 1996; Fuj ita et a!., 1999, 2000;
Fukamichi and Fujita, 2000), and the other is an isotropic large volume
expansion due to the magnetic-field-induced FOT. Figure 7. 43 demonstrates
the magnetic field dependence of the relative change in the length b. L / L II
parallel to the applied magnetic field direction at 5 and 50 K for LuCCo o.90
Gao. 10) 2 with T c -- 30 K CFig. 7. 10) CSaito and Fukamichi, 2001). This
compound exhibits 10- 6 orders of magnitude at 5 K. However, it should be
354 Kazuaki Fukamichi

noted that the value at 50 K above the Curie temperature is much larger than
that at 5 K because of the volume expansion caused by the metamagnetic
transition in the paramagnetic temperature range (Saito and Fukamichi,
2001).
6

~
x 3 LU(CoO.90GaO.10)2
'-'

~" Tc=30 K
<l 2

o 2 4 6 8 10
B(T)

Figure 7. 43 Magnetic field dependence of the relative change in the length llL / L II
parallel to the applied magnetic field direction at 5 and 50 K for Lu (COO. 90 Gao 10 )2
ferromagnetic compound with T c = 30 K (Saito and Fukamichi, 2001).

Next, we discuss the data for La ( Fe,- x Six) 13 having a higher Curie
temperature. To elucidate the magnitude of the volume change by the IEMT,
the linear magnetostriction measured parallel to the magnetic field direction,
I:.L / L II , just above the Curie temperature for x = O. 12 and O. 14 is shown in
Fig. 7.44 (Fujita et al., 1999; Fukamichi and Fujita, 2000). The change in
the length at the IEMT temperature is significantly large. The evaluated volume
change defined as three times of I:.L / L II at 9T is 1. 5 % and O. 9 % for x =
O. 12 and O. 14, respectively. The obtained result for x = O. 12 is comparable
with the volume change at the Curie temperature confirmed by X-ray diffraction
as given in Figs. 7.35 and 7.36. Therefore, the volume change defined above
should be close to the net volume change. By using the relation, I:. V / V =
KC rnv M2 , the values of KC rnv are obtained as about 7 x 10- 3 118 2 and 5 x
10- 3 118 2 for x = 0.12 and 0.14, respectively, being the same in magnitude
as those of Co-based Laves-phase compounds (Table 7. 1). Therefore, such a
significantly large volume change is due to the large magnetic moment induced
by the IEMT. Strictly speaking, the volume change in the magnetic field should
be obtained from the change in the length measured parallel and perpendicular
to the magnetic field. However, it is worth noting that the magnitude of three
times these values is very close to the spontaneous volume change due to the
FOT at T c (Fig. 7. 36). Therefore, it is concluded that the free energy as a
Itinerant-Electron Metamagnetism 355

function of both M and V has two local minima of a finite magnetic moment
with a large volume and null magnetic moment with a small volume.
0.6

x=0.12

0.4

o x=0.14
o
0.2 220 K

o 2 4 6 8 10
B (T)

Figure 7. 44 Linear magnetostriclion measured parallel to the magnetic field direction


ilL / L II just above the Curie temperature for La (Fe,- x Six) 13 with x = O. 12 and O. 14
(Fujita, et aI., 1999; Fukamichi and Fujita, 2000).

10
B(T)

Figure 7. 45 Change in the length ilL! L II at 200 K parallel to the appl ied magnetic field
for La(Feo.88 Sio. 12) '3' together with that of Fe1.Q05 (Rhu5 PdO. 15 )0.995 (Ibarra et aI., 1995)
and Fe70. Pd29 .6 (Furuya et aI., 1998). The room temperature anisotropic linear
magnetostriction, A II - A J-' that is, the difference between the linear magnetostriction
parallel (All) and perpendicular OJ-) to the magnetic field direction for a polycrystalline
TbFe2 (Clark, 1980) and the data at 265 K for Nb MnGa (Ullakko et aI., 1996).
356 Kazuaki Fukamichi

Figure 7.45 shows the change in the length t:.L / L II at 200 K parallel to
the applied magnetic field for La (FeOS8 Si o 12) 13 (Fukamichi, 1998; Fukamichi
and Fujita, 2000; Fujita and Fukamichi, 2001a), together with that of
Fe1005 (Rhos 5Pdo 15 )0.995 having an FOT from the antiferromagnetic to the
ferromagnetic state (Ibarra et aI., 1995) and Fe704Pd296 shape-memory alloy
(Furuya et al., 1998). The measuring temperature is also given in the same
figure. The room temperature anisotropic linear magnetostriction, A II - A1- ,
that is, the difference between the linear magnetostriction parallel (j\ II) and
perpendicular (A 1-) to the magnetic field of a polycrystalline TbFez (Clark,
1980), and the data at 265 K for Niz MnGa shape-memory alloy Wllakko, et aI.,
1996), are also shown in the same figure for comparison. The value of t:.L/
L II for La (Fe088 Si o 12 ) 13 greatly exceeds the value of A II - A 1- for the
polycrystalline TbFe2 compound, though the measured temperature is lower
than room temperature RT. As is well known, TbFe2 based magnetostrictive
materials exhibit a large All - A1-' while the volume change All + 2A1- is
nearly zero because the magnetostrictive properties are anisotropic. These
anisotropic magnetostrictive properties of TbFe2 are mainly originated from the
rotation of the localized magnetic moment of Tb. On the other hand, the
volume magnetostriction of the IEMT is due to the onset of the magnetic
moment caused by the exchange splitting of 3d electron bands. The
polarization of 3d electron bands causes a volume change proportional to the
square of the local magnetic moment, and both A II - A1- and A II + 2A 1- have
the finite values. A volume change of 1. 2 % sharply occurs at To in
La(Feo88 Sio 12) and the value of three times t:.L / L II is very close to that of the
volume change at To. Such an isotropic giant magnetostrictive property is
very attractive, compared with those of TbFe2 based anisotropic
magnetostrictive materials, because the giant magnetostriction of La (Fe088
Si O. 12 ) can be obtained even in a polycrystalline state without any
crystallographic structure controls. The field sensitivity of t:.L / L II in low
magnetic fields of La(Feo 88 Si o 12) is similar to that of TbFe2' An important
point to note is that the magnitude of Be of the IEMT becomes lower as the
temperature comes close to To = 195 K and it becomes zero at To, showing
the spontaneous volume change at To caused by the thermally induced FOT.
The value of Be extrapolated to room temperature for x = O. 12 exceeds a few
ten tesla, which is a relatively high field for conventional magnets. To realize
a large volume change in low transition fields around room temperature, it is
necessary to make the specimen with the higher transition temperature T 01 ,
because the transition field just above To becomes very low. The linear
magnetostriction parallel to the magnetic field direction for La (Fe088 Si o 12) HlO
is shown in Fig. 7.46 (Fujieda et aI., 2001). By applying magnetic field, a
marked change of the linear magnetostriction is observed around room
temperature. The magnitude of the linear magnetostriction t:.L / L II
corresponds to the value of the discontinuous linear thermal expansion
Itinerant-Electron Metamagnetism 357

h.Ws CT c) /3 due to the disappearance of the local magnetic moment given in


Fig. 7. 41. This indicates that the applying magnetic field just below T c
restores the ampl itude of the local magnetic moment CFuj ita et aI., 2002c).
The temperature dependence of the linear magnetostriction is correlated to the
temperature dependence of Be. Accordingly, such a large linear
magnetostriction around room temperature is attributed to the giant volume
magnetostriction of about 1% by the IEM transition. The magnitude of a linear
magnetostriction in the magnetic field of 7 T is about O. 3 %, larger than the
room temperature value of TbFerbased compounds. With decreasing
temperature, a large value is obtainable in lower magnetic fields as seen from
the figure. Therefore, such a giant magnetostriction is expected between T c
and 284 K in much lower applied magnetic fields, because Be of the IEMT
tends to become zero at T c .
0.35
La (Feo 88SiO 12)IJH 10
0.30 Tc=278K

0.25

0.20

~
--'I~ 0.15
~ -a-- 284K
0.10 --+- 286 K
-lr- 288 K
0.05

0 2 4 6 8
B(T)

Figure 7.46 Magnetic field dependence of the change in the length fJ.L / L II C== the linear
magnetostriction) parallel to the applied magnetic field direction just above the Curie
temperature T c for LaCFe088 Sio 12) 13 HIO CFujieda et aI., 2001).

For conventional linear magnetostrictive materials, crystallographic


structure controls are necessary because the magnitude of the linear
magnetostrictions exhibit a remarkable anisotropy against the crystallographic
axis. For ferromagnetic shape-memory alloys, crystallographic structure
controls are required in order to regulate the variants in the martensitic
transformations for the purpose of easy displacements of twin boundaries. In
contrast to these materials, it should be emphasized that La (Fe,- x Six) 13 Hy
requires no crystallographic structure controls. Consequently, the
LaCFel- x Six) 13 Hy system is promising as new-type giant magnetostrictive
materials.
358 Kazuaki Fukamichi

7. 8. 2 Giant Magnetocaloric Effect


Conventional gas compression technology for refrigeration is approaching its
limitation in efficiency. In addition, environmental problems caused by
refrigerant gases are serious. In such circumstances, recently, it has been
desired to develop magnetic refrigerants for cooling from room temperature to
temperatures as low as possible by using superconducting and permanent
magnets. An FaT such as the IEMT should theoretically take place at a
constant temperature, and hence the differential of M with respect to T, which
is the measure of the magnetic entropy change ~Smag, may become infinitely
large. Therefore, we might expect a giant magnetocaloric effect CMCE) in
LaCFel-xSix)13Hy' In this section, the recent results intended for high
temperature applications covering room temperature are discussed.
Comprehensive data for various other materials can be found in reviews of
Tishin C1999) and Pecharsky and Gschneider C1999) .
Given in Fig. 7. 47 are the magnetization curves above the Curie
temperature of 195 K as a function of magnetic field and temperature for La
CFe088SioI2)13CFujita et aI., 2002b). The solid lines trace the magnetization
curves in ascending magnetic field, except for the data obtained at 196 K just
above T c, in which descending magnetic field process is also shown in the
same figure. It should be noted that La CFeOS8 Sio 12) 13 exhibits the IEMT from
the paramagnetic to the ferromagnetic state by applying magnetic field.
Therefore, S-shape characteristic magnetization curves are observed in wide
temperature ranges. In ITMs, SFs in the paramagnetic state are strongly
enhanced by the exchange interaction, whereas those in the ferromagnetic
state are suppressed due to the onset of local magnetic moments. For
LaCFel-xSU13Hy, the change of the magnetic moment due to the IEMT
exceeds 1.5 IJs' and hence a remarkable magnetic entropy change would be
also correlated to the suppression of SFs. The magnetic entropy change ~Smag
is given by
(7.34 )

From the Maxwell relationship, ~S mag under the magnetic field change from HI
to H 2 is obtained from the differential of magnetization M with respect to T as

(dM)
f
2
~Smag = H
HI ar H dH. (7.35)

As shown by the dashed curves in Fig. 7. 47, the thermal sweep of


magnetization under the constant magnetic field exhibits a decrease in
temperature due to the increase of Be. The temperature of the inflection point
in the M vs T curve significantly increases with the increase of the magnetic
field, different from the SOT from the ferromagnetic to the paramagnetic
Itinerant-Electron Metamagnetism 359

2.0

2 B(T)
200 205
T(K)

Figure 7. 47 Variation of magnetization M as a function of magnetic field Band


temperature T above the Curie temperature T c for La (Fe088 Si o. 12 ) 13 (Fujita et aI.,
2002b) .

state. Therefore, a large ilS mag is expected in wide temperature ranges with
increasing magnetic field.
Shown in Fig. 7. 48 is the temperature dependence of ilS mag calculated
from the magnetization curves by using Eq. (7.34) in the field change of 0 - 1,
0-2 and 0 - 5 T for La (Feos8 Sio 12) 13 (Fuyieda et aI., 2002). A remarkable
maximum value of ilS mag about - 18 J/kg K in the field change of 0 - 1 T is
observed around T e of 195 K. It should be noted that La(Feos8 Sio 12) 13 shows
the thermal-induced FOT at T e, corresponding to the IEM at Be = O.
Therefore, ilS mag is large even in low magnetic fields. With increasing
magnetic field, the magnitude of the maximum value of ilS mag gradually
increases and a relatively large ilS mag is observed in wide paramagnetic
temperature ranges. Note that the values in the field change of 0 - 1, 0 - 2 and
0-5 T for La(Fe088Si012)13 are much larger than those of LaFe11.4Si16(Hu
et aI., 2001 b) because the former specimen contains no impurity phase such
as ex-Fe.
The Curie temperature T e of La ( Fe 1- x Six) 13 is easily controlled by
hydrogen absorption as seen from Fig. 7. 40. By adjusting the hydrogen
concentration, T e is continuously controlled up to 336 K (Fujieda et al.,
2001). Furthermore, the IEMT above T e occurs up to above room temperature
as seen from Fig. 7. 42. The temperature dependence of ilS mag for
La(Fe08SSio12)13Hl0 with T e = 283 K is presented in Fig. 7.49 (Fujita et aI.,
2002a). The peak of ilS mag vs T curve is observed around T e in the field
change of 0 - 1 T in a similar manner as that of the specimen without
hydrogen, although the maximum value becomes slightly small. With
increasing magnetic field, the ilS mag vs T curve exhibits a higher maximum
which is broad in shape, resulting in a wide plateau of - 21 J/ (kg K) around
280 - 290 K in the field change of 0 - 5 T. The closed and open circles in the
field change of 0 - 2 T show the data of the ascending and descending
processes of the magnetization curves, respectively. At every temperature,
360 Kazuaki Fukamichi

o
DO'DOC-D
pO'''
",11
11'
fl'
-10
~
,j
: 0-5 T
<I

-30 L - L- L -_ _---JL-_ _-----'


190 200 220 230

Figure 7.48 Temperature dependence of the isothermal magnetic entropy change ll.Smag
for La(FeoaaSi o12 ),3 as a function of magnetic field change (Fujita et aI., 2002)

-10

-30 L - '-- '-- --'

270 280 290 300


T(K)
Figure 7.49 Temperature dependence of the isothermal magnetic entropy change ll.Smag
for La(Feo.aaSio12)13Hl.O as a function of magnetic field change (Fujita et aI., 2002a)

the influence of hysteresis on the ilS mag curve due to the FOT is larger than
experimental error. From the ergodic nature of the phase transition, this result
corresponds to the small hysteresis of ilS mag against the thermal cycle.
Actually, a small hysteresis was observed in the thermomagnetization curves
measured with increasing and decreasing temperatures. From a practical
viewpoint, a large value and a plateau together with a small thermal hysteresis
Itinerant-Electron Metamagnetism 361

in the ~Smag vs T curves of the La(Feos8SioI2)13Hy are favorable


characteristics for magnetic refrigerants. The magnetic refrigerators require a
large MCE in a wide temperature range (Pecharsky et aI., 1999). Fe49 Rh s1
(Nikitin et aI., 1990; Annaorazov et aI., 1996) and Gds (Six Gel- x) 4
(Pecharsky and Gschneidner, 1997, 1999) have been considered to be good
candidates for the magnetic refrigerants in the vicinity of room temperature. In
these materials, a characteristic first-order phase transition occurs around
room temperature and the transition above the transition temperature is
controlled by magnetic field. However, a large irreversibility in the phase
transition prevents these materials from applying to magnetic refrigerants. In
more detail, the magnetocaloric effect in Fe49 Rh s1 disappears after several
thermal cycles in an alternating magnetic field (Annaorazov et aI., 1996). For
Gds(Ge xSi 1- x) 4' a martensitic structural change at the transition temperature
causes a gradual change of the transition temperature after thermal cycles
(Levin et aI., 2001; Choe et aI., 2000). In contrast to these materials, it
should be emphasized that the magnetic phase transition in La(Feos8Sio12)13Hy
is practically reversible without any structural transformations.
By changing hydrogen concentration, T c can be controlled from 195 to
336 K. The value of ~Smag at T c in the field change of 0 - 5 T as a function of
hydrogen concentration for La(Feo88SioI2)13Hy is plotted in Fig. 7.50 (Fujita
et aI., 2002a). It seems that the hydrogen absorption has a very small
influence on the magnitude of ~Smag. As seen in Fig. 7. 50, the specimen
where y = 0 shows the peak in the ~Smag vs T curve under 0 - 5 T, and the
-15

1.0

Y~
-25 !-;,-;,---::-!-::-----:::-':.".--------;:-!:-::------,~
180 210 240 270 300
reCK)

Figure 7.50 Magnetic entropy change .t.S mag in the field change of 0 - 5 T at the Curie
temperature T c as a function of hydrogen concentration y for La(FeoaaSio.12),3Hy (Fujita
et al., 2002a).

peak temperature, 197 - 210 K, with a plateau of 21 - 23 J/ (kg K) are


similar to each other after hydrogen absorption. Therefore, ~Smag in magnetic
fields in the paramagnetic temperature region is very similar in every specimen
with a different hydrogen concentration. In Fig. 7.51, the value of ~Smag in the
field change of 0 - 5 T against temperature for La ( Feo 88 Sio 12 ) 13 H1 0 is
362 Kazuaki Fukamichi

compared with the data reported for Fe systems such as FeO.49 Rhosl
(Annaorazov et aI., 1996) and La(Fe094CoO 06)1183A1117 (Hu et aI., 2001a).
For comparison, the value in the field change of 0 - 5 T for Gd (Benford and
Brown, 1981) taken for the reference is given in Fig. 7.52, together with the
data of Gds(Si OS Ge05)4 (Pecharsky and Gschneidner, 1997). The value of
La(Feo88Si012)13Hl0 is about twice that of Feo.49 Rho.sl , La(Fe094Coo06)1183AI117
and Gd. Furthermore, the curve of La(Fe088SioI2)13HlO becomes a plateau
keeping a large value, whereas that ofGds (Si 05 Ge05)4 exhibits a sharp peak
in the Iimited temperature range. The adiabatic temperature change I:!. Tad is
given by the following expression:
(7.36)

-5

.-. -10
Q La (FeO.94CoO.06)11.8JAI1.I7 FeO.49RhO.51
On
.0<
"'::'

~
-15

-20 \
La (Fe088Siol2)1JHLO
0-5T
-25
200 250 300 350
T(K)
Figure 7.51 Temperature dependence of the magnetic entropy change /:;.8 m in the field
change (/:;.H) of 0 - 5 T for La(Feo.88 Si o. 12 ) 13 H1.o (Fujita et al., 2002a), together with that
of FeO.49 Rho.51 (Annaorazov et aI., 1996), La(Fe09,Co006)II8JAI117 ( Hu et aI., 2001a).

The peak width of I:!.Smag is required to be affected strongly by applying


magnetic field, otherwise I:!. Tad is not so large even though I:!.Smag is
significantly large (Pecharsky and Gschneidner, 2001). The value of I:!. Tad is
evaluated using the temperature dependence of I:!.Smag and the specific heat
data. The temperature dependence of I:!. Tad for La(Fel-xSix)13 with x =0.88,
0.89 and O. 90 under the field change of 0 - 2 T is presented in Fig. 7. 53
(Fuj ieda et aI., 2002). With decreasing concentration x, the value of I:!. Tad
becomes larger. Figure 7. 54 shows the temperature dependence of the
adiabatic temperature change I:!. Tad in the field change of 0 - 2 T for materials
which exhibit a large value of I:!. Tad around room temperature. The temperature
dependence of the value of I:!. Tad for La(Fe088Sio12)13HlO is very similar to that
of Gds(Si 05 Geo s) 4(Pecharsky and Gschneidner, 1997) and its magnitude is
Itinerant-Electron Metamagnetism 363

-5
-- .... .......
~
"-

Q
bIl
-'"
-10
..,
'::::"

~
Gds(Sio.sGeO.s)4

-IS

-20 '--- ---"- -L -----.J

200 250 300 350


T(K)

Figure 7.52 Temperature dependence of the magnetic entropy change ~Sm in the field
change (~H) of 0 - 5 T for Gd (Benford and Brown, 1981) and Gds(Sio.sGeo.s)4
(Pecharsky and Gschneidner, 1997).

10

x=O.IO La(Fe'_xSix )13


8 0-2 T

6
g ."
~
<1 4

0
170 180 190 200 210 220
T(K)

Figure 7.53 Temperature dependence of the adiabatic temperature change ~ Tad in the
magnetic change of 0 - 2 T for La(Fel-x Six) 13 (Fujieda et ai., 2002).

large, comparable with that of La (Fe088 Sio 12) 13 and larger than that of MnAs
(Wada and Tanabe, 2001) as seen from figure. Listed in Table 7.3 are the
magnetic entropy change ~ S mag and the adiabatic temperature change ~ Tad
around the transition temperature T r under the field change from 0 to 2 T for Gd
(Benford and Brown, 1981), Gds (Si o.s GeO.S)4 (Pecharsky and Gschneidner,
1997), MnAs (Wada and Tanabe, 2001), La (Feo. 94 Coo 06 ) 1183 AI 117 (Hu et ai.,
364 Kazuaki Fukamichi

8
0-2 T

MnAs

OL- --'==-_ _----' ---'---_~_ .......


180 220 260 300 340
T(K)
Figure 7.54 Temperature dependence of the adiabatic temperature change /:!,. Tad in the
magnetic change from 0 to 2 T for Gds (Si 05 Geos), (Pecharsky and Gschneidner Jr., 1997>,
MnAs (Wada and Tanabe 2002), La(Feos8 Si o 12) 13 and for La(Feos8 Si o. 12 ) 13 H,o (Fujita et
aI., 2002a, Fuj ieda et aI., 2002).

2001a)and La(FeosBSioI2)13HlO(Fujita et aI., 2002a; Fujieda et aI., 2002).


It should be noted that the magnitude of 11 Tad does not correspond to that of
I1S mag (Pecharsky and Gschneidner, 2001). The value of 11 Tad for La (FeO.BB
Sio. 12 ) 13 HI. 0 is larger than that of Gd and MnAs. Although the value of 11 Tad for
Gds(SiQ5Geos)4 is very large, the crystallographic phase stability is not so
good (Levin et aI., 2001; Choe et aI., 2000), as mentioned in connection with
Fig. 7.49. On the other hand, by changing Fe concentration x, we can expect
a larger value of I1T ad for La(FexSi-x)13Hy in a similar way as Fig. 7.53.
Accordingly, La (FexSi- x )13 Hy can be regarded as promising refrigerants
working in a wide temperature range covering room temperature under
relatively low magnetic fields generated by permanent magnets.

Table 7. 3 The magnetic entropy change /:!,.$mag, the adiabaitic temperature change /:!,. Tad
and the transition temperature T, under the field change of 0 - 2 T for Gd (Benford and
Brown, 1981>, Gds (Si o s Geo s ), (Pecharsky and Gschneidner, 1997>, MnAs (Wada and
Tanabe, 2001), La ( Feo 9' Coo 06 ) 1183 AI1.l7 (Hu et aI., 2001) and La( FeOS8 Si o 12) 13 H, a
(Fujita et aI., 2002a; Fujieda et al. 2002).

Materials - /:!,. $ mag (J/ (kg' K Tr(K)

Gd 5 5.7 294 (1)


Gds (Sio.sGeos), 14 7.3 278(2)
MnAs 30 4.7 318(2)
La(Feo.9, COo. 06 ) 11.83 AI"7 5 273(1)
La(Feo. 88 Si o.12 ) 13 Hl.O 19 6.2 274(1)
( 1) Curie temperature;
(2) Structural transformation temperature.
Itinerant-Electron Metamagnetism 365

7.8.3 Giant Magnetoresistance


Giant magnetoresistance (GMR) effects have extensively been investigated
from practical and theoretical viewpoints. The reader is referred to the article
by Batzen and Grunberger in this series. The GMR effects in metamagnetic
compounds are briefly presented in this section. The magnetoresistivity of
itinerant-electron magnets is written as
li.p ( B, n = li.pc ( B , n + Ii.PSF ( B , n (7.37)

where Ii.Pc is the positive contribution due to the cyclotron motion of conduction
electrons and Ii.PSF is the negative contribution due to spin fluctuations. In the
case of the IEMs, the latter term Ii.PSF would become dominant because the
spin fluctuations are significant as explained in the preceding sections.
Figure 7.55 depicts the magnetoresistance effect parallel to the magnetic field
direction at 5 K for Lu(Coo912AI0088)2(Yokoyama and Fukamichi, 2001). The
arrows exhibit the increasing and decreasing processes of applying magnetic
fields. An obvious hysteresis is caused by the IEMT, and a marked decrease
takes place around Be. The negative large magnetoresistance is about 13 %
in 8 T (Yokoyama and Fukamichi, 2001), similar to the GMR in magnitude of
superlattices as well as granular alloys. Between Lu (COO 912 A1 0088 )2 in the
paramagnetic state and Lu ( Coo 902 Al o 098 ) 2in the ferromagnetic state, in this
context, the difference in the electrical resistivity P at 4. 2 K is less than
2

-2

-4
~

e -6
~
<I
-8

-10

-12

-14
0 2 4 6 8 10 12
B (T)

Figure 7.55 Magnetic field dependence of the magnetoresistivity parallel to the applied
magnetic field direction at 5 K for Lu (Coo 912 Al o.oaa )2 (Yokoyama and Fukamichi, 200 1) ,
the value of p(O) is the resistivity in zero field and I:!.p is the difference in resistivity after
applying magnetic field.
366 Kazuaki Fukamichi

1 %, which is negligibly small. Therefore, such a GMR is not attributed to the


change in the electronic state from the paramagnetic to the ferromagnetic
state, but dominantly to the suppression of SFs. The reason for a slight
increase in the lower fields remains as a matter to be discussed further. The
magnetoresistance perpendicular to the magnetic field direction also shows a
similar tendency. Shown in Fig. 7. 56 is the temperature dependence of the
magnetoresistance effect perpendicular to the applied magnetic field direction
for ferromagnetic Lu ( Coo 902 Al o 098 ) 2 (Yokoyama and Fukamichi, 2001). The
decrease of the magnetoresistivity is remarkable ~n the vicinity of the Curie
temperature T c, showing a clear suppression of SFs by applied magnetic
field.
2

o
- - IT
-0- 2T
---- 3 T
--{}- 4 T
-- 5T
-0- 6T

-4 ......... 7 T
-6- 8 T

--><-- 9 T

-6 l------,l,---------,L----:'::----:':-----,-J
o W ~ W W 100
T(K)
Figure 7.56 Temperature dependence of the magnetoresistivity I:1p/ p(5K) perpendicular
to the applied magnetic field direction for ferromagnetic Lu(C00902 Al oo98 )2 (Yokoyama and
Fukamichi, 2001).

From these data, it is expected that La (Fel- x Six) 13 metamagnetic


compounds would also be accompanied by a GMR. However, at the present
stage, microcracks caused by the large volume expansion due to the IEMT
prevent us from measuring the magnetoresistance.

7.9 Concluding Remarks

Metamagnetic materials are fascinating because the metamagnetic transition


exhibits drastic changes in various magnetic and electrical properties. For
research and development, it will be useful to keep the following points in
Itinerant-Electron Metamagnetism 367

mind.
Since the IEMT occurs in the limited concentration range as given in
Figs. 7.2 and 7. 25, and the transition is closely correlated with spin
fluctuations, homogeneity in composition is very important for fundamental and
practical researches. By using specimens with a compositional fluctuations,
that is, in an inhomogeneous state, almost all magnetic and electrical
properties such as Be' T c ' X max ( n, Yexp are different from their intrinsic
values (Yokoyama et aI., 1998). Therefore, it is important to keep in mind
that many results reported previously are different from the data given in this
chapter.
The itinerant-electron metamagnetism is sensitive to the peculiar band
structure in the vicinity of E F as illustrated in Fig. 7.4. In order to increase the
Curie temperature T c ' we tried a substitution of Fe by Co in La(Fe x Si 1 - x )13.
In La (Fe088 COy Sio 12- y ) 13 compounds, the Curie temperature T c linearly
increases and reaches 250 K at y = O. 04 (Fuj ita and Fukamichi 1999;
Fukamichi and Fujita, 2000). However, achievement of a large
magnetostriction at low fields was not so easy. Therefore, Co is not so helpful
as the substitutional element. This reason may be correlated with the
deformation of a pecul iar band structures of 3d electrons. Accordingly, deep-
rooted considerations for the electronic states are necessary to design IEMs for
practical applications.

References
Aharony, A. Critical Phenomena. In: F. J. W. Hahne ed. Lecture Notes in
Physics, No. 186. Springer, pp.209-258(1983)
Aleksandyan V. V., A. S. Lagutin, R. Z. Levitin, A. S. Markosyan and V.
V. Snegirev. JETP. 62: 153 (1985)
Aoki, M. and H. Yamada. J. Magn. Magn. Mater. 78: 377(1989)
Annaorazov, M. P. A., S. A. Nikitin, A. L. Tyurin, K. A. Asatyan and A.
K. Dovletov. J. Appl. Phys. 79: 1689( 1996)
Benford S. H. and G. V. Brown. J. Appl. Phys. 52: 2110 ( 1981)
Choe, W., V. K. Pecharsky, A. O. Pecharsky, K. A. Gschneidner Jr., V.
G. Young Jr. and G. J. Miller. Phys. Rev. Lett. 84: 4617(2000)
Clark, A. E. In: E. P. Wohlfarth, ed. Ferromagnetism, Vol. 1. North-
Holland, pp. 531 - 589( 1980)
Cyrot, M. and M. Lavagna. J. Appl. Phys. 50: 2333 ( 1979)
Duc, N. H. and P. E. Brommer. In: K. H. J. Buschow ed. Handbook of
Magnetic Materials, Vol. 12. Elsevier Science B. V., pp. 259 - 394 (1999)
Duc, N. H. and T. Goto. In: K. A. Gschneidner Jr. and L. Eying. eds.
Handbook on the Physics Chemistry of Rare Earths, 26. Elsevier Science
B. V., p. 177 - 264 ( 1999)
Entel P. and M. Schroter. Physica B 161: 160 (1989)
Fujieda, S., A. Fujita, K. Fukamichi, Y. Yamazaki and Y. lijma. Appl.
368 Kazuaki Fukamichi

Phys. Lett. 79: 653 (2001)


Fujieda, S., A. Fujita and K. Fukamichi. Appl. Phys. Lett. 81: 1276(2002)
Fujita, A., Y. Akamatsu and K. Fukamichi. J. Appl. Phys. 85: 4756(1999)
Fujita, A. and K. Fukamichi. IEEE Trans. Magn. 35: 3796 (1999)
Fujita A., K. Fukamichi, F. Gejima, R. Kainuma and K. Ishida. Appl. Phys.
Lett. 77: 3054(2000)
Fujita, A. and K. Fukamichi. J. Magn. Soc. Jpn. 25: 66(2001a)
Fujita, A. and K. Fukamichi, unpublished (2001b)
Fujita A., S. Fujieda, K. Fukamichi, Y. Mitamura and T. Goto. Phys. Rev.
B 65: 014,410(2002a)
Fujita A., S. Fujieda, K. Fukamichi, Y. Yamasaki and Y. lijima. Mater
Trans. 43: 1202(2002b)
Fujita, A., K. Fukamichi, J. T. Wang, and Y. Kawazoe. Phys. Rev. B 68:
10, 441 (2003)
Fujita A., S. Fujieda, K. Fukamichi, Y. Yamasaki and Y. Iijima. Trans.
Mater. Research. Soc. Jan. 26: 219(2001)
Fukamichi, K. In: F. E. Luborsky ed. Amorphous Metall ic Alloys.
Butterworths Co. Ltd, London, p. 317 ( 1983)
Fukamichi, K., In: W. Gorzkowski, H. K. Lachowicz and H. Szymczak eds.
Proc. the 4 th Int. Cont. on Phys. Magn. Mater. World Sci. Pub. Co.,
Singapore, p. 354 (1989)
Fukamichi, K. J. Magn. Soc. Jpn. 22: 173(1998)
Fukamichi, K. and A. Fujita. J. Mater. Sci. Technol. 16: 167(2000)
Fukamichi, K., T. Yokoyama, H. Saito, T. Goto, H. Yamada. Phys. Rev.
B 64: 134,40 1( 200 1)
Furuya, Y., N. W. Hagood, H. Kimura and T. Watanabe. Mater. Trans. JIM
12: 1248( 1998)
Goto, T., H. Aruga-Katori, T. Sakakibara, H. Mitamura, K. Fukamichi and
K. Murata. J. Appl. Phys. 76: 6682(1994)
Goto, T., K. Fukamichi, T. Sakakibara and H. Komatsu. Solid State
Commun. 72: 945( 1989)
Goto, T., T. Sakakibara, K. Murata and K. Fukamichi. J. Magn. Magn.
Mater. 90&91: 700(1990)
Goto, T., Y. Shindo, Y. Takahashi, S. Ogawa. Phys. Rev. B 56: 14,019
( 1997)
Goto, T. and M. I. Bartashevich. J. Phys. : Condens. Matter 10: 3625
(1998)
Goto T., K. Fukamichi and H. Yamada. Physica B 300: 167(2001)
Hayashi, K., K. Tajima, H. SaitoandK. Fukamichi. J. Phys. Soc. Jpn. 69:
4013(2000)
Hu, F. X., B. G. Shen, J. R. Sun and Z. H. Cheng. Phys. Rev. B 64: 012,
409(2001a)
Hu, F. X., B. G. Shen, J. R. Sun, Z. H. Cheng, G. H. Rao and X. X.
Zhang. Appl. Phys. Lett. 78: 3675(2001b)
Itinerant-Electron Metamagnetism 369

Ibarra, M. R., P. A. Algarabel, C. Marquina, Y. Otani, S. Yuasa and H.


Miyajima. J. Magn. Magn. Mater. 140/145: 231(1995)
Iijima, M., K. Endo, T. Sakakibara, T. Goto. J. Phys. : Condens. Matter
2: 10,069 (1990)
Ikeda, K., S. K. Dhar, M. Yoshizawa and K. A. Gschneidner, Jr. J. Magn.
Magn. Mater. 100: 292 (1991)
Irisawa, K., A. Fujita and K. Fukamichi. J. Alloys Compds. 305: 17(2000)
Ishiyama, K., A. Shinogi and K. Endo. J. Phys. Soc. Jpn. 53: 2456 (1984)
Klimker, H., M. P. Dariel and M. Rosen. Phys. Chem. Solids 40: 195
(1979)
Levin, E. M., A. O. Pecharsky, V. K. Pecharsky and K. A. Gschneidner,
Jr. Phys. Rev. B 63: 64,426(2001)
Mignot, J. M., J. Flouquet, P. Haen, F. Lapierre, L. Puech and J. Voiron.
J. Magn. Magn. Mater. 76/77: 97(1988)
Mohn, P. and K. Schwarz. J. Magn. Magn. Mater. 104/107: 2067(1992)
Moriya, T., Fluctuations in Itinerant Electron Magnetism. Springer- Verlag,
Berlin (1985)
Mariya, T. J. Phys. Soc. Jpn. 55: 357(1986)
Muraoka, Y., M. Shiga and Y. Nakamura. J. Phys. Soc. Jpn. 42: 2067
(1979)
Murata, K., K. Fukamichi, H. Komastu, T. Sakakibara and T. Gato. J.
Phys. : Candens . Matter 3: 2515 ( 1991)
Murata, K., K. Fukamichi, T. Sakakibara T. Gota, H. Aruga-Katari. J.
Phys. : Condens. Matter 5: 2583 ( 1993a)
Murata, K., K. Fukamichi, T. Sakakibara, T. Gata and K. Suzuki. J.
Phys. : Candens . Matter 5: 1525 (1993b)
Murata, K., K. Fukamichi, T. Gato, H. Aruga-Katori, T. Sakakibara and K.
Suzuki. Physica B 201: 147 (1994a)
Murata, K., K. Fukamichi, T. Gota, K. Suzuki and T. Sakakibara. J.
Phys. : Condens. Matter 6: 6659 ( 1994b)
Mushnikov, T. Goto, K. Kamishima, H. Yamada, A. V. Andreev, Y.
Shiokawa, A. Iwaa, V. Sechovsky. Phys. Rev. B 59: 6677 (1999)
Nikitin, S. A., A. G. Myalikglev, M. Tishin, M. P. Annaarazov, K. A.
Astryan and A. L. Tyurin. Phys. Lett. A 148: 363(1990)
Nishihara, Y. and Y. Yamaguchi. J. Phys. Soc. Jpn. 53: 2201( 1984)
Palstra, T. T. M., J. A. Mydosh, G. J. Nieuwenhuys, A. M. van der Kraan
and K. H. J. Buschaw. J. Magn. Magn. Mater. 36: 290( 1983)
Palstra, T. T. M., G. J. Nieuwenhuys, J. A. Mydash and K. H. J.
Buschow. Phys. Rev. B 31: 4622 (1984)
Pardew, J. P., K. Burke and M. Ernzerhaf. Phys. Rev. Lett. 77: 3685
(1996)
Pecharsky, V. K. and K. A. Gschneidner, Jr. Appl. Phys. Lett. 70: 3299
( 1997)
Pecharsky, V. K. and K. A. Gschneidner, Jr. J. Magn. Magn. Mater. 200:
370 Kazuaki Fukamichi

44(1999)
Pecharsky, V. K. and K. A. Gschneidner, Jr. J. Appl. Phys. 90: 4614
(2001)
Saito, H, T. Yokoyama, K. Fukamichi, H. Mitamura and T. Goto. J. Phys. :
Condens. Matter 9: 9333( 1997)
Saito, H., T. Yokoyama, K. Fukamichi, K. Kamishima, H. Mitamura and T.
Goto. Rev. High-Pressure Sci. Technol. 7: 556(1998)
Saito, H., T. Yokoyama, K. Fukamichi, K. Kamishima and T. Goto. Phys.
Rev. B 59: 8725 ( 1999)
Saito, H., T. Yokoyama, Y. Terada, K. Fukamichi, H. Mitamura, T. Goto.
Solid State Commun. 113: 447 (2000)
Saito, H. and K. Fukamichi. unpublished (2001)
Sakakibara, T., T. Goto, K. Yoshimura, M. Shiga, Y. Nakamura and K.
Fukamichi. J. Magn. Magn. Mater. 70: 126 (1987)
Sakakibara, T., T. Goto, K. Yoshimura and K. Fukamichi. J. Phys.:
Condens. Matter 2: 3381 (1990a)
Sakakibara, T., T. Goto, K. Yoshimura, K. Murata and K. Fukamichi. J.
Magn. Magn. Mater. 90/91: 131(1990b)
Shimizu, M. Proc. Phys. Soc. 86: 147( 1965)
Shimizu, M. Rep. Prog. Phys. 44: 329(1981)
Shimizu, M. J. Phys. (Paris) 43: 15( 1982)
Sumiyama, K., M. Shiga and Y. Nakamura. J. Phys. Soc. Jpn. 40: 996
(1976)
Sumiyama, K., Y. Emoto, M. Shiga and Y. Nakamura. J. Phys. Soc. Jpn.
50: 3296 ( 1981)
Takahashi, Y. J. Phys.: Condens. Matter 2: 8405 ( 1990)
Terao, K., and H. Yamada. J. Phys. Soc. Jpn. 66: 1063 (1997)
Tishin, A. M., In: K. H. J. Buschow ed. Handbook of Magnetic Materials.
North-Holland, Elsevier Sci. B. V., Amsterdam, pp. 395 - 524 ( 1999)
Ullakko, K., J. K. Huang, C. Kantner, R. C. 0' Handley and V. V.
Kokorin. Appl. Phys. Lett. 69: 1966(1996)
Wada, H. and Y. Tanabe. Appl. Phys. Lett. 79: 3302(2001)
Wagner, D., E. P. Wohlfarth. J. Phys. F: Mett. Phys. 11: 2417 (1981)
Wassermann, E. F. , In: K. H. J. Buschow and E. P. Wohlfarth ed.
Ferromagnetic Materials, Vol. 5. North Holland : Elsevier Sci. Pub.,
p. 238-321(1990)
Wohlfarth, E. P. and P. Rhodes. Phil. Mag. 7: 1817(1962)
Wohlfarth, E. P. J. Phys. C 2: 68(1969)
Yamada, H., J. Inoue, K. Terao, S. Kanda and M. Shimizu. J. Phys. F 14:
Met. Phys, 1943( 1984)
Yamada, H., J. Inoue and M. Shimizu. J. Phys. F : Met. Phys. 15: 169
(1985)
Yamada, H. and M. Schimizu. J. Phys. F 15: Met. Phys. L175(1985)
Yamada, H., T. TohyamaandM. Shimizu. J. Phys. F: Met. Phys. 17: L163
Itinerant-Electron Metamagnetism 371

( 1987)
Yamada, H. J. Phys.: Condens. Matter 3: 4115(1991)
Yamada, H. Phys. Rev. 847: 11,211(1993)
Yamada, H. and K. Terao. J. Phys.: Condens. Matter 6: 10,805(1994)
Yamada, H. J. Magn. Magn. Mater. 139: 162( 1995)
Yamada. H. Physica 8 211: 161( 1995b)
Yamada, H. Phys. Rev. 855: 8596( 1997)
Yamada, H. K. Fukamichi and T. Goto. Rhys. Rev. 865: 024,413(2002)
Yoshimura, K., K. Fukamichi, H. Yasuoka and M. Mekata. J. Phys. Soc.
Jpn. 56: 3652( 1987)
Yoshimura, K., Y. Yoshimoto, M. Mekata, K. Fukamichi and H. Yasuoka.
J. Phys. Soc. Jpn. 57: 2651(1988)
Yokoyama, T., H. Nakagima, H. Saito, K. Fukamichi, H. Mitamura and T.
Goto. J. Alloys Compds. 266: 13 ( 1998)
Yokoyama, T., H. Saito, K. Fukamichi, K. Kamishima, H. Mitamura and T.
Goto. J. Magn. Soc. Jpn. 23: 442(1999)
Yokoyama, T. and K. Fukamichi. unpublished(2001)
Yokoyama, T., H. Saito, K. Fukamichi, K. Kamishima, T. Goto and H.
Yamada. J. Phys.: Condens. Matter 13: 9281(2001)

The author wishes to thank Professor T. Goto of the Institute for Solid State Physics, the
University of Tokyo, Professor H. Yamada of Faculty of Science, Shinshu University,
Professor K. Tajima of the Department of Physics, Faculty of Science and Technology, Keio
University, Professor Y. lijima, Drs. A. Fujita, H. Saito, T. Yokoyama and K. Murata of
the Department of Materials Science, Graduate School of Engineering, Tohoku University for
collaborations and helpful discussions. Financial supports of Grant-in-Aid for Scientific
Research (6) (2), No. 13555168, and on Priority Areas (A) No.299 from Japan Society for
the Promotion of Science, and Industrial Technology Research Grant Program in 00
(00A26019a) from the New Energy and Industrial Technology Development Organization
(NEDO) of Japan are gratefully acknowledged.
8 Modeling of Hysteresis in Magnetic Materials

D.C. Jiles, X. Fang and W. Zhang

8. 1 Introduction

It is very useful if the magnetic behavior of a material can be precisely


described in advance, given the composition, structure and processing
history. Not only can the design cycle based on "trial and error" iteration be
reduced and replaced with materials design from first principles, but also
materials with higher performance can be achieved at a lower cost and with
less effort if materials scientists are able to employ viable computer-aided
design (CAD) software modeling tools.
It is in the nature of materials science and engineering that materials
properties depend on many factors at an extremely complicated level that is
usually beyond the scope of what theories based on one or two equations can
describe. Even though progress is still being made to make materials CAD
simulators more applicable and practical, much more is still needed. Before a
material with the desired performance can be designed, an accurate model is
needed, no matter whether it is empirical or theoretical, as long it can provide
enough predictive accuracy. For magnetic materials, a model accurately
describing hysteresis loops should be the foundation of such CAD tools. Many
magnetic properties such as coercivity, remanence, susceptibility and energy
loss can then be determined from the modeled hysteresis loop. This will not
only give scientists a better tool to probe the physics behind the magnetization
processes taking place inside a material, but such a model will also benefit the
permanent magnet, soft magnetic materials, data storage and sensors
industries by making materials by design truly feasible.
There were many efforts in the past to develop an analytical model that
fits actual experimental magnetization data. Models have been suggested to
describe the magnetization of ferromagnetic materials. However, only some of
these have had prolonged success. These include the Preisach model
(Preisach, 1935), the Stoner-Wolhfarth model (Stoner and Wolhfarth, 1991),
the Globus model (Globus, 1975), and the Jiles-Atherton model (1986). A
detailed review of these models has been given by Liorzou et al. (2000).
Before these models can be used in a meaningful way, one has to be aware of
the strengths and weaknesses of each model, since none of these is capable of
Modeling of Hysteresis in Magnetic Materials 373

giving a truly comprehensive description of the many magnetic processes that


take place in a magnetic material. A direct comparison of these models is
given in Table 8. 1.

Table 8.1 Comparison of four hysteresis models after Liorzou et al. (2000) .
Characteristic Preisach Stoner-Wohlhfarth Globus Jiles-Atherton

Mechanism Not specified Rotation Wall motion Not specified


Anisotropy Not specified Uniaxial Multi-aixs Multi-aixs
Interaction Moving model Yes No Yes
Pinning Not specified Yes Yes Yes
Anisotropic or Uniaxial Anisotropic or
Texture Not specified
isotropic (180 wall) isotropic
Wall energy No No Yes No
Additional
Reversibility Additional model Yes Yes
model
Minor loops Yes Yes Approximation
Demagnetization Yes Yes
Anhysteretic Yes Yes Yes Yes

Parameters M s' P(h.,h b ) , M s' K' , O' M s' y' M s ' e*


, o'
ex ex D' , f' a*, k*
Grains Not specified Single domain Bi-domain Multi domain
Computational
++++ +++ + ++
time

Magnetic Soft magnetic Bulk materials


Hard magnetic
Materials recording materials - Medium
materials
Thin films mainly ferrites ferrites
* Measurable parameters.

8. 2 Development of Model Theories of Hysteresis

Hysteresis is a commonly-occurring phenomenon in nature arising most often as


a result of cooperative behavior of a large number of identical interacting
elements (Vicsek, 2001). The most familiar examples occur in
ferromagnetism, but similar behavior occurs in ferroelectrics (Chen and
Montgomery, 1980), ferroelastics (Tuszynski et aI., 1988) and fatigue. For a
long time, theory and modeling of hysteresis in magnetic materials was a
subject for the specialist investigator. However, in recent years the
widespread and increasing capability and accessibility of computers has made
modeling available to a much wider range of investigators, so that this area of
374 D.C. Jiles, X. Fang and W. Zhang

study has become of much wider interest.


There are a number of reasons for the high level of interest in the subject
of modeling the magnetic properties. The first, of course, comes from the
continuing need to develop new and improved magnetic materials for the ever-
increasing range of applications in which they are used. In order to improve
materials on a systematic scientific basis, it is necessary to have a deep
understanding of the underlying physics of magnetization and, in particular, the
relationship between the structure and properties of magnetic materials. This
understanding can then be used to direct research along paths that are likely to
be successful in identifying materials with the desired properties. Another
reason for modeling the magnetic properties is in the design of devices, using
computer aided design methods, whereby the performance of materials with
different properties can be assessed and compared. These techniques can be
used to identify suitable, and even optimal, materials properties for particular
applications and can show how devices are likely to perform. This approach
means that much of the expense in designing and constructing prototype
devices can be eliminated through the use 'of computer-aided modeling
techniques so that design changes and optimization are performed on a
computer and the final construction is then based on the ideal design and an
accurate description of the magnetic properties.
The magnetic modeling methods that are in use today are quite diverse,
and the choice of model depends crucially on the length scale of interest.
Ultimately, these models depend on the presence of magnetization (magnetic
moment per unit volume) in certain materials. This magnetic moment per unit
volume can be represented in terms of a net magnetic moment per atom,
although in many cases, but not all, the magnetic moment is not actually
localized on the atomic/ionic cores, but instead is caused by itinerant
electrons. To understand how this net magnetic moment comes about, theories
were developed by Stoner ( 1933) and Slater ( 1936) concerning the electron
band structure of magnetic materials, which explained why there is an
imbalance of spin-up and spin-down electrons in ferromagnets. Band structure
calculations provide a description of magnetism at a very fundamental level
and much work is still being performed to refine and develop band structure
calculations (Antropov et aI., 1996). However, although these provide a basis
for understanding magnetism on the scale of individual electrons and atoms,
such theories are not central to this paper because they are not used for
hysteresis modeling of materials.
The models that we will consider here range from first principles
calculations at a level of discrete magnetic moments, such as those of Landau,
Lifschitz, and Gilbert, to methods that can be used to model whole magnetic
components of devices on the macroscopic scale, such as the Preisach model.
Modeling of Hysteresis in Magnetic Materials 375

8. 3 Magnetism at the Discrete Level of Individual


Atoms and Beyond to the Continuum Level: Landau-
Lifschitz-Gilbert Model and Micromagnetics

The principal idea of the Landau-Lifschitz-Gilbert model is to describe the


behavior of individual discrete magnetic moments under the action of a
magnetic field. Clearly, from the classical times of Ampere, it was known that
on the macroscopic scale a magnetic dipole moment of fixed magnitude would
rotate under a magnetic field according to the equation

T = JJo m + H, (8.1)

where T is the time constant, JJo is the permeability of free space, m is the
magnetic moment, and H is the magnetic field.
This same concept, the general torque equation, was applied at the
atomistic scale, or more correctly at the level of discrete magnetic moments of
fixed magnitude but variable orientation, by Landau and Lifschitz (1935). This
approach can be used to describe the behavior of an individual magnetic
moment. The behavior of the entire material can then be determined by
integrating the same process over the entire solid. The rate of change of
magnetization with time then depends on the torque
aM
at =- Yr T (8.2)

where M is the magnetization, t is the time, and Y r is the gyromagnetic ratio.


In the absence of damping a magnetic moment that is not initially aligned with
the total field will precess around the field direction with a resonance frequency
<'<>0' which depends on the strength of the field and the magnitude of the
moment,

<'<>0 = YrJ.1 o H (8.3)

where Yr is the gyromagnetic ratio of the magnetic moment. In the complete


absence of damping, this precession will continue for infinite time as shown in
Fig. 8. la. In practice there is always some damping in solids and therefore for
light damping (long time constant) there will be some precessional motion and
some rotational motion towards the field direction as shown in Fig. 8. 1b. The
time taken to do this will depend on the damping coefficient. At high levels of
damping (short time constant) the precessional component of the motion is
suppressed because the moment reaches its final equilibrium orientation before
precession has taken place, as shown in Fig. 8. lc.
Laying aside for the time being the practical problems of actually
376 D.C. Jiles, X. Fang and W. Zhang

z z z

t t t
H H H
(a) (b) (c)

Figure 8. 1 The motion of magnetic moment under the action of a magnetic field depends
crucially on the level of damping. (a) no damping results in precession; (b) light damping
results in a spiralling motion of the magnetic moment into the field direction; (c) heavy
damping results in a rotation of the moment in to the field direction with no precession.

performing the calculations for such a large number of individual magnetic


moments (typically there are 1028 atoms per unit volume in a material such as
iron, each carrying it own moment), there are other considerations that must
come in to play. The above equations describe the behavior of magnetic
moments in isolation. In a magnetic material there are interactions between
the moments, so that magnetism in sol ids is a cooperative process. Therefore
the above equation must be modified when dealing with moments in a solid to
take into account these interactions. Landau and Lifschitz suggested the
following modification,

3M =_ _ 4ITA2 d(M x M x H) (8.4)


3t YT M

where the Ad is the damping coefficient, the second term on the right hand side
of the equation is a damping term which restrains the rotation of the moments
under the action of a field. The existence of this damping term was, and
remains, a hypothesis. There is no direct experimental verification, but the
existence of such a term in bulk materials is reasonable from a theoretical
standpoint.
A crucial question arises over the exact form of the magnetic field H that
should be used in this equation. For an isolated magnetic moment in free
space, this will be just the classical magnetic field obtained from Maxwell's
equations. However, in view of the magnetic interactions that exist within a
material this field must also include any other interactions which give rise to a
turning force on a dipole. An obvious and well-known example of this is the
exchange interaction, which is not included in the classical description of the
magnetic field under Maxwell's equations.
The damping coefficient in the model equation is treated as an adjustable
parameter, since there appears to be no obvious way to determine this from
Modeling of Hysteresis in Magnetic Materials 377

first principles. A value that is consistent with experimental data is used, and
this is then used to determine the behavior of an array consisting of a large
number of individual magnetic moments under the action of an applied field.
The model does not explicitly include temperature in the calculations, although
one can argue that the damping coefficient is temperature dependent.
However, this provides only one rather arbitrary way of including temperature
effects in the model.
Twenty years after the original paper by Landau and Lifschitz, Gilbert
wrote independently on the subject, from a continuum perspective, and
developed a modified form of the equation (Gilbert, 1955). The Gilbert form
of the equation, as shown below, is used mostly for model ing today.

dM
dt
=_ YlJo M x H + 41T Ad (M x d M) .
YlJo M 2 dt
(8.5)

In principle the Landau-Lifschitz-Gilbert (LLG) model can be used either


at the atomistic scale or at the "micromagnetic" continuum scale with the unit
cells being much smaller than a single domain. The model is really directed
towards volumes smaller than a single domain. Practical applications of the
Landau-Lifschitz equation are made at length scales much greater than the
atomistic. Typically, individual "moments" are defined at the level of 50 nm
and each of these moments is treated as an indivisible unit. This can be
justified if there is coherent rotation of the atomic moments occurring over the
volume occupied by these moments. The model is suitable for time dependent
calculations of the reorganization of magnetic moments under the action of a
field. It has the advantage over other models of not requiring the moments
within a domain to be parallel. In fact it calculates the optimum orientations of
moments, which in general are not necessarily parallel, as shown in Fig. 8.2.

i~i~@$;i{~;~i~~~1i~}1fi
... ---- ------------------------ -- -._-._--- --
,/ ....-. -"
-._--.-._-------------------~-~-.----._----~ ...... -... ... -
..
,/
J,
.....
,/ ,/ _
,/
_ __ ___ -.
I/-----_-~---------------------.------------~
,/ - - - - - - -
-.-'-"-
1,/,//-----------------------------_-
,/ ,/ ..
,/~~/~---------------------.--_._._._----~---,
,/ .,/ / . r - - __. __ ...- ..- _ .. .._ ... --
--' --'-- -- '-"""
..- - - - - " - - - - . - - - - - - - - - - - - - - - - - . - - -~ --~ --~ - - - - - - - - - - - -
------ .. --- -
.. ..
,
- - __ - - - - -
.-._-.~._--------

- - - - .... - - - - ... - -.- ...._ ... ,./ _ '\


-- -- -- - - - - - - -
-.~-
- , - . --- __-

~
"

'I
'I;
"':.

'\
.... ..,.//.:~.... ~:..:-..... :. - _ --_ ...... _t
--_ --.... - .. --.... -- --_ ..... - .... - _ -.. - - : --: ....-_ -:: ",:"",:,_-- "':;.._--.:- --- - - , / .,! ~ A t .....

Figure 8. 2 Solution of the Landau-Lifschitz-Gilbert equation for an array of magnetic


moments representing a thin film of permalloy of dimensions 2 IJm by 0.5 IJm. For the
purposes of calculation each individual magnetic moment occupies and area of O. 05 IJm x
0.05 IJm. The solution shows the configuration of magnetic moments is not all parallel.
378 D. C. Jiles, X. Fang and W. Zhang

The results of calculations based on the Landau-Lifschitz-Gilbert model


can be averaged to provide hysteresis curves of materials as shown in
Fig. 8.3. The model can be extended beyond the range of the single domain
into simulations of multi-domain specimens through the use of finite element
methods. This approach is known as "computational micromagnetics" and
enables both micro- and macro-scale calculations to be incorporated into a
single model simulation (Kronmuller et al., 1997; Fidler and Schrefl, 2000).
This allows material microstructure to be included in the simulation and in this
way it is possible to combine the advantage of exact theoretical equations
developed for a small number of interacting magnetic moments with
appl ications to practical materials.
1.0

0.5

---6-- my
-0.5 --..- mxdown
- - mxup
-1.04-_0lI!!!!!I!!1!!!I!!!~~l..-_----, _
-50 -25 25 50

8.4 Magnetism of Domain Rotation: Stoner-Wohlfarth

The Stoner-Wohlfarth model takes as its basis an array of single-domain


particles which can reorient their magnetization by coherent rotation of all
moments within the domain. As such it takes no account of the details of the
individual moments below the single domain in scale, and unlike the LLG
model, it does not try to account for the orientations within the domain.
The model addresses a dispersion of single domain magnetic particles in
some other unspecified matrix material. This is alluded to in the title of the
paper when it refers to "heterogeneous" alloys. On the other hand, an
assumption of the model is that the magnetic particles are all identical. Within
domains the model considers the competing effects of anisotropy and magnetic
field on the orientation of moments as shown in Fig. 8. 4. The domains
themselves can have random alignments or they can be textured (meaning
Modeling of Hysteresis in Magnetic Materials 379

preferred orientation) or they can be even completely aligned in certain


directions, although this would be a relatively trivial case requiring nothing
more than a calculation of moment orientations within one single domain
particle.

~='-..J.-------H

Figure 8. 4 Rotation of magnetization Ms of an isolated single domain under the


competing effects of magnetic field H and anisotropy H K

The model in its original formulation assumed that there were no magnetic
interactions between the particles. In other words, the distribution of the
particles was so dilute that each particle was effectively isolated and could not
be influenced by the orientation of any other particles. This assumption can be
changed, and has been changed by others, although including such interactions
adds greatly to the computational complexity of the model. The model in its
original condition also assumed axial anisotropy. This is the simplest type of
anisotropic calculation to make. Other forms of anisotropy such as cubic
anisotropy can be included (Lee and Bishop, 1966).
The basic idea of the model is to consider the reorientation of a magnetic
moment within a single domain particle in which the applied field is at some
arbitrary angle to the anisotropic easy axis. In the case of the uniaxial easy
axis along the field direction (Fig. 8. 5), this results in a bimodal switching

Ms

HK H

(a) (b)

Figure 8.5 Rotation of the direction of magnetization Ms of a single domain particle with
uniaxial anisotropy, oriented with its easy axis H K along the direction of the applied field H.
380 D.C. Jiles, X. Fang and W. Zhang

behavior with attendant coercive fields in the forward and reverse directions.
In the other extreme case, where the magnetic field is appl ied along the
anisotropic hard axis, this results is a magnetization curve with no coercivity
(Fig. 8. 6). In general the domains will be oriented at an arbitrary angle
relative to the field and such domains will have properties that lie between
these two extremes (Fig. 8. 7). A complete material may then consist of an
assembly of domains, each at different angles to the field direction (Fig. 8. 8).
The overall magnetization of the multidomain sample is then the vector sum of
all of the magnetic moments of the domains divided by the total volume as
shown in Fig. 8. 9.

H H

(a) (b)

Figure 8.6 Rotation of the direction of magnetization M s of a single domain particle with
uniaxial anisotropy, oriented with its easy axis H K perpendicular to the direction of the
applied field H.

HK M

Ms

K/Ms H

I
I
H
(a) (b)

Figure 8.7 Rotation of the direction of magnetization M s of a single domain particle with
uniaxial anisotropy, oriented with its easy axis H K at an arbitrary angle to the direction of
the applied field H.
Modeling of Hysteresis in Magnetic Materials 381

0.,10.,90
1.0
o 80 45

'::i;'"
~ 0 ------ ---
-s-
'"o
u

45 80 o
-1.0 0.,10.,90
-1.0 o 1.0
h

Figure 8.8 Magnetization curves obtained using the Stoner-Wohlfarth model for domains
with uniaxial anisotropy oriented with different directions of the easy axis relative to the
applied field.

1.0

0.5

'J 0
~
-0.5

-1.0
-1.5 -1.0 -0.5 0 0.5 1.0 1.5
H
Figure 8. 9 A composite hysteresis curve obtained by summing the elementary
magnetization curves for a randomly distributed orientation of easy axes.

The turning force on a magnetic moment due to the applied field depends
on the vector product of a magnetic moment with a magnetic field. The turning
force on the magnetic moment due to anisotropy is the derivative of the energy
with respect to angle. For a particle with uniaxial anisotropy the equilibrium
orientation is given by
/.10 HM s sin4> - 2Ksinecose =0 (8.6)
where 4> is the angle of the magnetization relative to the field, K is the
anisotropy coefficient and e
is the angle of the magnetization relative to the
magnetic easy axis. From the Stoner-Wohlfarth model it is possible to
calculate the saturation field H s needed to rotate the magnetization of the most
diffcult domain oriented at 90 to the field direction,

H =~ (8.7)
s /.10 M s

where H s is the saturating field. It is also possible to calculate the coercivity of


382 D. C. Jiles, X. Fang and W. Zhang

the domains based on the switching field needed to reorient the domain aligned
antiparallel to the field,
K
H =-- (8.8)
e IJo M s

where He is the coercive field. The model has been widely used for describing
the dependence of magnetic properties of materials on anisotropy and texture.
In addition, many of the original ideas behind the model have been developed
and extended and have found appl ications in fine particle systems (Spratt et
aI., 1988; Chantrell, 1997).

8. 5 Magnetism at the Level of Domain Boundaries:


Neel, Globus-Guyot, Bertotti

Although in most cases the boundaries of magnetic domains comprise only a


small fraction of the total volume of a magnetic material, in multi-domain
samples much of the magnetization change occurs in these regions, and
therefore it is of great interest to know and understand what changes are
occurring at the boundary in order to predict and model the properties of such
materials.
The importance of domain boundary motion has been recognized by many
investigators including Becker (1932, 1943), Kersten (1956), Neel (1944,
1959), Globus (Globus and Guyot, 1972, 1973; Globus and Duplex, 1966,
1969; Globus et aI., 1971), Escobar (Escobar et aI., 1983, 1985) and more
recently by Bertotti (Bertotti, 1988; Alessandro et aI., 1990). The main idea
in the treatment of domain boundary motion is to separate the motion into two
components: reversible and irreversible. In most respects this separation is
somewhat artificial, since both processes take place together in multi-domain
materials. However, the physics of the situation can be more easily analyzed
if these processes are separated. Irreversible processes necessarily cause
energy dissipation and lead to coercivity and hysteresis while reversible
processes do not.
The motion of domain walls can further be divided into two types: flexible
wall motion and rigid wall motion. The higher the surface energy of the domain
walls, the greater tendency to move in planar fashion, particularly if the
pinning sites that restrain the domain wall are weaker. On the other hand,
domain walls with lower surface energy will tend to bend first, particularly if
the pinning sites are stronger. In practice components of both types of
movements occur together.
In the case of planar wall motion, much of the early modeling was
developed by Becker who was particularly concerned with the movement of
Modeling of Hysteresis in Magnetic Materials 383

domain walls interacting with regions of inhomogenous strain, such as


dislocations. Simplifying assumptions need to be made in order for an
analytical solution to be obtained. Assuming a simplified periodically varying
internal potential through which the domain walls move, an expression for the
0
susceptibility for a material composed of 180 domain walls can be derived in
terms of the amplitude and wavelength of the periodic potential,

_ 2AI 2 J.10 M~COS2 e (8.9)


X- 3lT2AsO"o

where X is the susceptibility, e


is the angle of the domain magnetization
relative to the applied magnetic field, A is the cross sectional area of the
wall, 1 is the periodicity of the internal potential with amplitude E rnax = 3As 0"0 /2
due to periodically varying stress of amplitude 0"0 in a material of
magnetostriction As. The critical field, or local coercivity occurs when the
domain wall moves over one of the local maxima in the slope of the potential
and is therefore given by,

He = EmaxlT (8.10)
1J.10 Mscose

Models have been developed by Neel and Kersten in the case of domain
wall bending as shown in Fig. 8. 10. Kersten in particular was concerned with
the movement of domain walls that were pinned by "inclusions," which meant
simply inhomogeneous volumes within the material.

I
I
I

Mst ~--_:: /
/
/
/
/

Figure 8. 10 Deformation of a domain wall under the action of an applied magnetic field

In this case equations can be derived for the initial susceptibility Xin for
small deformation of the domain wall
_ J.10 M~ 1 h
3

Xin - 3y (8.11)

where 1 is the spacing between the pinning sites and h is the length of the
section of domain wall as shown. For the local coercivity or critical field He
above which the domain wall will break away from the pinning site
384 D.C. Jiles, X. Fang and W. Zhang

He = ycOSeperit C8.12)
lJoMs/Ccos8 1 - cos8 2 )
where 8 1 and 82 are the angles of the magnetization relative to the magnetic
field on either side of the domain wall.
In the case of ferrites in which the defects are mostly confined to grain
boundaries, the domain walls will be pinned principally at the grain boundaries
and therefore the behavior can be modeled assuming that the domain walls
move simply by bending like an elastic membrane as described by Globus
C1962). The Globus model depends on equations for deformation of the domain
walls similar to those of Kersten. For the purposes of modeling, the grains are
assumed to be spherical. Most applications of ferrites are for higher-frequency
magnetic field, and therefore the domain walls can be considered to vibrate
under the action of a time-dependent field.
Under the action of a "weak" applied magnetic field the domain walls
deform but remain fixed on the grain boundaries. From this it was predicted
that the permeability depended linearly on the grain diameter. Comparison
with experiment yielded confirmation CGlobus and Duplex, 1966). Further
studies show that wall motion components dominate in ferrites, while rotational
processes, which are dependent on anisotropy but not grain size, are of
secondary importance in these materials CGlobus and Duplex, 1969). The
equations governing the deformation of the magnetic domain walls is described
by Globus et al. CGlobus et al., 1971). Although the model was developed
originally to describe the properties of ferrites, it was later shown to be valid
for spinels and garnets. Dissipative processes resulting from translation of the
domain walls at higher field strengths were added to the model subsequently
by Globus and Guyot C1972, 1973).
The most comprehensive treatment of the underlying theory of this model,
including most of the equations, which were not given in the papers by Globus,
has been given by Escobar et al. C1983, 1985). Accordingly the initial
susceptibility due to domain wall bending is
2
X = _Mys 0_ C8.13)

where M s is saturation magnetization, y is the domain wall surface energy and


o is the grain diameter. The value of the critical field for the unpinning of
domain walls is
2f
He = Mso C8.14)

where f is the pinning force per unit length on the domain wall along the grain
boundary, which is assumed to be independent of the applied field.
Bertotti et al. have extended some of the domain wall modeling concepts
through the use of domain wall eddy current dissipation and stochastic process
models in which they considered the domain wall as moving through a randomly
Modeling of Hysteresis in Magnetic Materials 385

fluctuating potential (Alessandro et a!., 1990; Bertotti et a!., 1999). The


randomness of the potential seen by the domain walls can be quantified using
two independent parameters, which mathematically represent the physical
properties of the potential in terms of an average amplitude and an average
fluctuation wavelength.
If He is a randomly fluctuating function of position, then it will also be a
randomly fluctuating function of the flux 4>. Local maxima in He represent
physical pinning sites in the material, and for large displacements of domain
walls, there will be a correlation length S which represents the range of
interaction of domain walls with pinning sites. These effects can be described
if it is assumed that the local coercivity He obeys a Langevin equation of the
form

dH e = He - (He) = dW (8. 15)


d<P s' d<P

where now the flux <P is the measure of displacement of domain walls instead
of the position, and W ( 4 is a randomly fluctuation function (or "white noise"
function), whose average value will be zero, and s' is the interaction length,
or correlation length, for domain walls with pinning sites. Eventually with some
additional restricting assumptions the equation reduces to

d<D + J..- (<D _ J,Jo AM) = __1_ dH e . (8. 16)


dt T aG dt

The motion of domain walls through the internal randomly fluctuating


potential leads to discontinuous changes in magnetization. These discontinuous
processes are manifested as Barkhausen noise (Schlesinger, 2001), which is
closely connected with the existence of hysteresis (Bertotti, 1996).
Hysteresis is found to be a direct result of discontinuous, dissipative processes
occurring over small volumes which, when summed together, produces the
familiar hysteresis over larger volumes. The Barkhausen emissions are fractal
in nature which means that the structure of the Barkhausen emissions is
independent of the scale. The underlying physics of these complicated types of
processes are only now being uncovered and discussed (Bertotti, 1998). A
recent review of the theory behind domain wall dynamics has been given by
Jiles (2000).
The domain boundary models are best applied to multi domain materials in
which the movement of magnetic domain boundaries is the principal
magnetization mechanism. This refers to mostly bulk soft magnetic materials
with low anisotropy and with large density of inhomogeneities either in the form
of strains (dislocations) or inclusions (particles of a second phase) .
386 D.C. Jiles, X. Fang and W. Zhang

8. 6 Magnetism at the Macroscopic Scale: the Integration


of Single Domain Switching Processes and the
Preisach Model

The Preisach model is a general mathematical model which describes


hysteresis on the macroscopic scale. The model was first developed to treat
magnetic hysteresis but the mathematical structure is equally applicable to
other physical systems exhibiting hysteresis, such as ferroelectric or
mechanical hysteresis. The model treats magnetic hysteresis as simply a
summation of a large number of microscopic switching events occurring in a
magnetic material. It was actually a development of previous work on
hysteresis by Weiss and Freudenreich and also derives some of its ideas from
the earlier model of Ising.
The Preisach model describes materials as an array of domains, each
with the same magnetic moment per unit volume (magnetization), but with
different switching fields as shown in Fig. 8. 11. The allowed microstate of the
magnetic moment is either "up" or "down," as in the Ising model. This spin-up
and spin-down restriction does limit the relevance of the model to actual
physical reality in most magnetic materials. The domains in the Preisach
model remain the same size, but there is no fundamental problem with this
approach for empirical modeling since the volume fractions of the domains with
particular combinations of switching fields can be varied within the model.
I I
b
i
i
i
i H
I
i
_ a--i---.!!---
i
i
i
Figure 8. 11 In the Preisach model each domain is represented by its saturation
magnetization and two characteristic switching field strengths h a and h b which describe the
field strengths needed to change the direction of magnetization.

The magnetic characteristics of a material are represented as the volume


fraction of domains with particular combinations of the switching field. This is
described by the probability distribution function P (h a , h b ) over the Preisach
plane defined by all possible combinations of h a and h b , which is the span of
Modeling of Hysteresis in Magnetic Materials 387

values of the two switching fields. The Preisach plane is shown in Fig. 8.12.
The probability density function Calso known as the Preisach function) P varies
over the span of possible values of switching fields and this represents the
distribution of different types of domain in the material.
y A y A' y

+ -
-
-
-
// -
/" " " A"
+ - +
+
- + - +
+ ~
-
-:t" - a
+-
+ + - +
+ + + - +
+ + + + -
+++++ B
y'
(a) (c)

y Alii y y
A'

O"{' ~~ -,, -__+-:


-H 0'" + + a

0"
I +++C;:
t--:t
t ~ =+ 0" +- + + t:
t1,,~++t
t B'
+ +'*: + + - B'"
Y" B" y' B ft
B'"
(d) (e) (f)

Figure 8. 12 The range of all possible pairs of switching fields Ch a , h b ) defines the
Preisach plane as shown here. The characteristics of a particular material are represented
as a probability distribution function PC h a , h b ) on the Preisach plane.

The magnetization M of the system as a function of applied field H Ct) can


be calculated by

MCt) = M s II
ha~hb
PCha,hb)oCha,hb)HCt)dhadhb C8.17)

where M s is saturation magnetization. The value of 0 Ch a , h b ) is either + 1 or


- 1 depending on the magnetic history.
The Preisach function can also be expressed as a function of he = Ch a -
h b ) /2 and h m = Ch a + h b ) /2, where he is the coercive field of an elementary
hysteresis loop and h m is the displacement or offset of the center of the
hysteresis loop from H = o. This representation of the Preisach function,
PC he' h m), allows an integrated coercive field distribution P Che) to be
computed by

C8.18)

where P Che) represents a probability distribution function for elementary


hysteresis loops having a particular value of coercive field he in the system.
Calculation of the Preisach function in order to describe the magnetic
properties of the material is therefore central to the use of the Preisach model
388 D.C. Jiles, X. Fang and W. Zhang

since it determines how many domains, or what volume fraction of the


material, will change orientation of its magnetization between two values of
the magnetic field. The Preisach function itself is not an invariant. It can
change with exposure to different field histories.
The description of the model first appeared in a restricted form which has
come to be known as the classical Preisach model. The classical Preisach
model has some characteristic properties, the most significant of which are
<D the memory (or wiping out property) whereby only the alternate series of
dominant field exposure maxima are remembered and (2) the congruency
property, whereby all minor hysteresis loops corresponding to the cycling of
the magnetic field between the same two extrema are congruent. The first of
these characteristic properties seems to be universally applicable to hysteresis
phenomena. The second is more limited in its applicability, and there are
certainly many instances where it is not valid.
The switching fields, or local coercivities, are assumed to be different
from one domain to the next. Furthermore within this model the switching fields
(coercivities) can be different in the up and down directions for the same
domain, although the model offers no explanation for this. Originally there
were no interactions allowed between the domains. Again this restriction is
rather unrealistic since there is unquestionably an exchange interaction within
magnetic materials, although the strength of this interaction between domains
is crucially dependent on domain size. A comparison of the classical Preisach
model to the measured magnetic hysteresis loops is shown in Fig. 8. 13, from

S' 0.04 S' 0.04


E E
~ 0.02 ~ 0.02
<:: <::
o o
.~ 0 .~ 0
.~ .~

Sb-0.02 Sb -0.02
oj oj

:::E -0.04 ~",==~e~~_--,L--_-,---1- :::E -0.04


Lk~====-------l..-_,.L--,--~L-,-
-2000 -1000 0 1000 2000 -2000 -1000 0 I000 2000
Mag. field (Oe) Mag. field (Oe)

S' 0.04 S' 0.04


E E
~ 0.02 ~ 0.02
<:: <::
o o
.~ 0 .~ 0
.~
!-0.02 Sb -0.02
oj oj

:::E -0.04 :::E -0.04


l.-.b~::r:::::=~_--l.--------l_ L-b==~--..L_--..L_--..L_
-2000 -1000 0 1000 2000 -2000 -1000 0 1000 2000
Mag. field (Oe) Mag. field (Oe)

Figure 8. 13 Comparison of measured and modeled hysteresis loops using the classical
Preisach model (Mayergoyz, 1991).
Modeling of Hysteresis in Magnetic Materials 389

which it can be seen that the model does broadly describe hysteresis, but
systematic discrepancies occur.
Generalizations of the original concept have been developed to address a
wider range of magnetic hysteresis phenomena (Mayergoyz, 1991). The
generalized Preisach model does not have the congruency restriction. These
later developments of the model also included interactions between domains,
but once this is included the actual locations within the material of domains with
particular orientations become significant. One way to circumvent the latter
problem is to use a mean field approach to model the interactions. This is
equivalent to assuming that each domain in the material interacts equally with
all other domains within the material, although this is unlikely to be realized in
practice. An example of the comparison of the generalized Preisach model
calculations to experimental results is shown in Fig. 8. 14.

0.04 0.04 x Classical


S' S' Experimental
E
~ 0.02 ! 0.02 results
c: c:
o .9
.~ 0 r1 0
.~ .~
c:
~ -0.02 ~-0.02
~ ~
-0.04 -0.04
-2000 -1000 0 1000 2000 -2000 -1000 0 1000 2000
Mag. field (Oe) Mag. field (Oe)

0.04 x Classical 0.04 x Classical


S' Experimental S' Experimental
!c: 0.02 results !c: 0.02 results
.9 .9
~ 0 .~ 0
.~ Q)
~ -0.02 ~ -0.02
~ ~
-0.04 -0.04
~~===:::::~~=----=-~
-2000 -1000 0 1000 2000 -2000' -1000 0 1000 2000
Mag. field (Oe) Mag. field (Oe)

Figure 8. 14 Comparison of measured and modeled hysteresis loops using the modified
Preisach model (Mayergoyz, 1991).

The quasi-static Preisach model describes the hysteresis losses in


materials. When the frequency of the exciting field is changed, however,
there are additional power losses, which are traditionally separated into
"classical losses" arising from the solution of the classical Maxwell's equations
in a magnetically permeable and electrically conducting medium, and" excess
losses". Dupre et al. (1998) investigated these frequency dependent power
390 D.C. Jiles, X. Fang and W. Zhang

losses using the dynamic Preisach model. It was known from experimental
observations that the excess losses, due to domain wall processes, depend on
the frequency of excitation according to an f312 power law and it was found that
this can be described under restricted conditions by the Preisach model.
Subsequently, Dupre et al. (1999a) investigated the description of power
losses using the generalized moving dynamic Preisach model under a
unidirectional applied magnetic field. Agreement was generally good over a
range of frequencies up to 1 kHz. Some systematic discrepancies were noted
at the higher end of this frequency range, which is probably attributable to the
influence of partial penetration of the magnetic field, due to the skin effect, on
the dynamic parameter obtained from the statistical theory of domain
processes. Furthermore, it was shown that the classical losses cannot be
derived through the statistical theory of domain processes.
The relationship between microstructure and magnetic properties is
important for understanding the behavior of magnetic materials. This can be
achieved through computational micromagnetics as described above, and as
shown by Dupre et al. (1999b) through the general ized dynamic Preisach
model. The effects of grain size and texture were separated within the model,
and the Preisach function was rewritten containing explicit terms representing
grain size and texture. These two terms were found to be the same for all
magnetization processes, including quasi-static reversible, quasi-static
irreversible and excess losses.
As shown by Pasquale et al. (1998); the Preisach and Jiles-Atherton
models are equivalent under certain conditions. In particular, if the Preisach
distribution function decreases exponentially with he according to the equations
above, then the rate of change of magnetization with field in the two models
can be shown to be equivalent.
In conclusion, the Preisach model provides a mathematical basis for
describing hysteresis. The model reconstructs the behavior in terms of a set of
more elementary components and has some similarities to the use of Fourier
analysis for periodic functions (I. Tomas private communication) in as much as
the total response (hysteresis) is represented as the sum of a large number of
elements (hysterons, each with different switching fields). The model is
widely applicable on the macroscopic scale for describing hysteresis in
magnetic materials, but the mathematics of the process give switching field
distributions that may not be physical. One severe limitation of the Preisach
model as formulated is that it really only allows irreversible processes to be
modeled. Reversible processes, if they are to be included must be taken care
of in an arbitrary manner. In addition, the basic Preisach model is scalar in
nature (due to the spin-up/ spin-down restriction), whereas magnetization
processes are in practice vector. The generalized Preisach model includes an
extension to describe the vector nature of these processes.
Modeling of Hysteresis in Magnetic Materials 391

8. 7 Magnetism at the Multidomain Level: Energy


Considerations and the Jiles-Atherton Model

In dealing with the behavior of materials, in particular their bulk magnetic


properties such as coercivity, remanence, permeability and hysteresis loss,
other problems arise that make it difficult if not impossible to simply scale up
the predictions of models that are based on consideration of one or two
domains. Therefore, a more general approach is needed in order to develop
equations that represent the average behavior of the materials. These models
necessarily use statistical thermodynamic principles to describe the resulting
magnetization behavior of a very large number of magnetic domains.
The earliest thermodynamic approaches were developed for the simplest
systems, specifically paramagnets. Paramagnets have the simpl icity of being
magnetically homogeneous, unl ike ferromagnets. Later models were
developed for the more technically important class of ferromagnets without
including hysteresis, and finally hysteresis models were developed. The model
of hysteresis developed by Jiles and Atherton depends on statistical mechanics
and is most relevant on the mesoscopic scale. It works well for materials with
low anisotropy for which the main mechanism is domain boundary movement.
It can be used for simple anisotropies such as axial and planar anisotropies
with minor modifications. For highly anisotropic materials, it can still be used
with the understanding that a simple analytic anhysteretic equation cannot in
general be developed for anisotropic materials, and therefore the
mathematical approximations become less realistic the greater the anisotropy
and the larger the number of magnetic easy directions.
The classical model for magnetism is the Langevin-Weiss model which
considers an array of magnetic moments in thermal equilibrium at a particular
temperature. This was used by Jiles and Atherton as the basis for developing a
model of hysteresis. The orientations of the magnetic moments are distributed
statistically, and integrating the distribution of moments over all possible
orientations leads to an equation for the bulk magnetization. The details of this
depend on the restrictions imposed by anisotropy, so that, for example,
different solutions are obtained depending on whether the magnetic moments
experience axial anisotropy, planar anisotropy or are in a completely isotropic
environment (Jiles et ai., 2000).
The extension of the Langevin-Weiss theory used to describe
ferromagnetic materials incorporates coupling among magnetic moments,
acting as a strong magnetic field to align the magnetic moments in a domain
parallel to each other. To quantify this coupling, a mean field which is
proportional to the bulk magnetization He = H + aM is used. This mean field
392 D.C. Jiles, X. Fang and W. Zhang

approach to describing the interactions needs to be applied with some caution,


but recent work by Chamberlin (2000) has shown that the mean field approach
is viable for clusters of spins on the nanoscopic scale.

8. 7. 1 Description of the Anhysteretic Magnetization


By replacing the classical magnetic field H with the effective magnetic field
H + aM, which inlcudes coupling to the magnetization, an equation for the
anhysteretic magnetization of a ferromagnetic material can be obtained as
follows:

M = M L [J.l 0 m (H + aM) ] (8. 19)


s kBT

where a is the mean field compl ing coefficient, and k B is Boltzmann's


constant. Instead of considering coupling between each individual magnetic
moment, the mean field is used to represent the inter-domain coupling.
For isotropic materials the anhysteretic function is
M = Ms[coth(x) - l/x] (8.20)

where x = J.l: Brr;.R, R = H + aM, and T is temperature.


Alternatively for materials exhibiting axial anisotropy,
M = Mstanh(x) (8.21)
and for materials exhibiting planar anisotropy

M = k B T ~Iogz =M I~(x) (8.22)


J.lo dH s 'o(x)'

where

'o(x) = ~ (5 1)2
1
(~ t (8.23)

and

, ~ 5 (X )2S-1 (8.24)
' 0 (x) = ;S (51)2 "2
A more generalized extension to cover other more complicated
anisotropies was made (Ramesh et al., 1996), in which the energy of a
magnetic moment with anisotropic perturbation was calculated in three
dimensions, and therefore different kinds of anisotropic materials could be
described. An increasing range of magnetic materials in which anisotropy and
texture playa significant role, for example hard magnetic materials, can be
seen in Section 8.3.
Modeling of Hysteresis in Magnetic Materials 393

Following the development of the generalized anhysteretic function (Jiles


et al., 1997)

2..: e- E1kB T cose


M s all moments
(8.25)
M aniso
2..: e- E1kB T

all moments

where e is the angle between the direction of the magnetic moment and the
direction of the applied field, and

E = IJ 0 < m >(H + aM) + E aniso (8.26)


where E aniso is the anisotropy energy which depends on the anisotropic
structure of material. For example, in the case of cubic anisotropy,
3

E aniso = K1 2..: COS


2
e COS e
i
2
j (8.27)
i*"i

with the normal convention on symbols. In this description, only the first
anisotropy coefficient K 1 was used since this approximation is in most cases
sufficient to provide an accurate description of the different magnetization
curves along different directions. A texture coefficient f text ' which is a
statistical evaluation of the volume fraction of the textured portion of a
material, was also introduced. The anhysteretic magnetization can then be
given as
Man = f text M aniso + (1 - f text ) M iso (8.28)
where Maniso is the anisotropic anhysteretic magnetization contribution, and Man
is the isotropic anhysteretic magnetization contribution. For more complicated
textured materials, there may be several different texture orientations such
that each particular direction has a proportion of the grains oriented along it. In
these cases the anisotropic contribution of each part must be calculated
separately and the net anisotropic portion of the anhysteretic magnetization is
the weighted sum of the components of magnetization of these orientations
along the direction of the applied field.
From this description of the thermodynamic anhysteretic magnetization, it
is possible to develop a description of hysteresis through consideration of
energy dissipation mechanisms. The irreversible and reversible components of
magnetization can be described separately in mathematics, although they are
linked physically. The two components of magnetization can then be combined
to give an equation for the total magnetization.

8.7.2 Extension to Describe Hysteresis


Under the action of a magnetic field, domain walls move so that the volume of
domains al ign favorably with respect to the field direction increases, at the
394 D.C. Jiles, X. Fang and W. Zhang

expense of the domains aligned unfavorably with respect to the field direction.
In the absence of energy dissipation all energy supplied to the material is equal
to the change in magnetostatic energy in the material. This is anhysteretic
magnetization. In the case of hysteresis, the energy suppl ied to the material
equals the magnetostatic energy plus hysteresis loss. The magnetostatic
energy in the material is the energy difference between input energy and the
energy loss due to processes such as domain wall pinning. One of the premises
of the model is that the energy loss is proportional to the change in
magnetization (Fig. 8. 15). Although this was derived from the domain wall
motion under the action of a magnetic field, it is not limited to domain wall
motion magnetization and, therefore, the model applies to any situation in
which the energy loss is proportional to the change in magnetization, as could
also occur under domain rotation.
An equation for the irreversible change in magnetization is obtained,

dMirr = __M--=an_-_M---,i,,--rr__ (8.29)


dH ok - a(M an - M irr )
where the directional parameter 0 takes + 1 when H increases in the positive
direction (dH/dt>O), and - 1 when H increases in the negative direction
(dH / d t < 0), ensuring that the pinning always opposes change in
magnetization. This differential equation represents the irreversible component
of magnetization.
During magnetization there is also a reversible component of
magnetization that can result from reversible domain wall bowing, reversible
translation of domain walls or reversible domain rotation. For the purposes of
modeling the reversible component M rav of magnetization was assumed to be
proportional to the difference between the anhysteretic magnetization Man and
irreversible magnetization M irr , with a constant of proportionality, the
reversibility coefficient e, which represents the fraction of magnetization
change that is reversible, so that 1- e represents the fraction of magnetization
change that is irreversible. Hence, the total magnetization M is the sum of
reversible magnetization and irreversible magnetization.
M = M irr + M rev = (1 - e) M irr + eM an (8.30)
where the constant coefficient e ranges from 0 (completely irreversible
magnetization) to 1 (completely reversible magnetization). The model
equation for the total magnetization, which includes both irreversible and
reversible magnetization, is then

dM (1 ) Man - M dMan (8.31)


dH = - e ok - a(M an - M) +e dH .

Solutions of this equation with a suitable anhysteretic function give a


typical sigmoid-shaped hysteresis loop. Through changing the parameters, this
model is able to predict the magnetization of soft and hard magnetic materials.
Modeling of Hysteresis in Magnetic Materials 395

400
- - Preisach
E 200 ----- Jiles-
~ Atheton
c:
o
.~ 0
.~
v
~
~ -200

-400 L---4L---_L - - - 0 L - - -L-----...JL - -


2 2 4
Magnetic field (kA/m)

Figure 8. 15 Comparison of measured and modeled hysteresis curves of cobalt modified


gamma iron oxide material (Andrei and Stancu, 1999)

8. 7. 3 Extension to Describe the Effects of Stress on Magnetization


The effects of stress on magnetization of materials can also produce very
significant changes. The incorporation of these effects into a more general
model, which also includes magnetic field and temperature, has been
achieved. The key to this is to provide a description under which both
magnetic field and stress can be treated as similar. An equation for the stress
equivalent field has been identified as (Sablik and Jiles, 1988; Sablik et aI.,
1987) ,

H =~~(~) (8.32)
u 2 J.1o J M T
where H u is the stress equivalent field, a is the stress, and i\ is the
magnetostriction.

Hu(e) = ~ ~ (cos2e - vSin2e)(:~)T (8.33)

where v is the Posson's ratio.


Variable stress also has effects that go beyond those that are described
by the above equation. In fact the application of stress causes unpinning of
domain walls, and this effect can be described by a law of approach to the
anhysteretic (Jiles, 1995; Jiles and Devine, 1995). This is given by the
following equation for the change in irreversible component of magnetization
M irr with elastic energy W:

dM irr
dW = I1 (M an -
M)
irr
(8.34)
396 D.C. Jiles, X. Fang and W. Zhang

and adding on the reversible component gives the change in the total
magnetization,

dM = + an
dW I1 CM an - M) C
dM
dW . C8 . 35)

The anhysteretic is itself stress dependent because of the contribution of


H q to the effective field, so the law of approach really contains two
components: a reversible component that represents the change in the
anhysteretic with stress, and an irreversible component which represents the
change in the displacement of the magnetization from the prevailing stress
dependent anhysteretic.

8. 7. 4 Extension to Describe the Effects of Frequency on


Magnetization
The effects of the frequency of the magnetic field on magnetization can also be
included in this model CJiles, 1994). In this case the effects of eddy currents
add to the dissipation, and so result in higher coercivity and hysteresis loss.
2
dH)(dM)2
IJ O d (GdWlJoHo)I/2(dH)I/2(dM)3/2
( 2p{3 dt dH + p dt dH

+ [k5 - a( ManCH) - M(H) + k5c ~~:n) J(~~)


- [ ManCH) - MCH) + k5c (~~:n) ] = 0 C8.36)

where d is the diameter, G is the dimensionless eddy current parameter, and


5 is the directional parameter.
Other approaches to frequency dependence of magnetization are needed
for insulating materials, in which eddy currents do not playa major role. This
has been described previously for the case of high frequency ferrites CJiles,
1993). The differential equation governing the magnetization in this case is

C8.37)

where TJ is the damping coefficient for harmonic motion and W n is the natural
resonance frequency of harmonic motion.
As can be seen from Table 8. 1 this model and its extensions, as given in
the above equations describe magnetic properties in terms of a multi-domain
structure that makes it widely applicable. The physical basis of this model is
an open, self-adjustable system to cover the case of anisotropic and textured
structures. This eliminates the limitation of the isotropic approximation and
expands the applicability of this model to more complicated situations.
In the model, the differential magnetic susceptibility depends on the
displacement of the prevailing magnetization from the anhysteretic
Modeling of Hysteresis in Magnetic Materials 397

magnetization. The anhysteretic magnetization is a function of the energy of


the moments within a domain. To include anisotropic effects into the model,
the anisotropy energy must be incorporated into the total energy of the
moments.
The general equation of hysteresis can be solved with the incorporation of
the appropriate anisotropic, textured or stress dependent anhysteretic
magnetization Man. This gives the magnetization curves along particular
directions. An advantage of this model is that the basic hysteresis equation
remains the same as do the hysteresis coefficients, so that the only change is
the incorporation of different forms of anisotropy into the equation for the
anhysteretic.
It should be mentioned here that the calculated resultant magnetization M
is the component parallel to the direction of the appl ied field, because the
anhysteretic and hysteresis equations represent the component of
magnetization along the field direction. As a result. although the model
equation for hysteresis remains basically the same, there is a significant
difference in the modeled magnetic properties along different field directions
due to the differences in the anhysteretic magnetization along different
directions.

8. 7. 5 Applications
During the early stage of the development of this model, efforts were directed
mainly towards iron-based materials. The results show that the model can
provide flexibility to describe magnetic properties of a wide range of iron
based alloys with carbon contents from 0 - 1 wt. %. and manganese zinc
ferrite. In addition, detai led equations for calculating parameters have been
given.
8.7.5. 1 Soft Magnetic Materials
The model has been used to simulate anisotropic magnetization curves of soft
magnetic materials such as iron, nickel and cobalt as shown in Figs. 8. 16.
8.17 and 8.18 (Jiles, 1997,1998). The model takes into account that the only
difference between the magnetization curves in the different crystallographic
directions is due to the anisotropy coefficient while. for a given material, all
properties and, hence, all other model coefficients remain the same. The
theoretical curves are in agreement with experimental observations. This
result shows that the hysteresis equations represent the underlying physical
processes, and once the parameters for a particular material have been
established they can be used as a sound basis for extending the scope of the
model description to other effects.
The model has also been used in conjunction with finite element methods
for solving periodic steady state magnetic field modeling problems. In this
398 D.C. Jiles, X. Fang and W. Zhang

1.0

0.8 Ms=505 kNm


a=575 Aim
::{ 0.6 a=0.0055
~ 0.4
- - Modeled
o Experiment
k=1 kA/m
c=0.6
K a=-4.5 kJ/m 3
0.2
Halong( II 0)
o
o 10 20 30 40 50
Appled magnetic filed (kNm)

1.0

0.8 Ms=505 kA/m


a=575 Aim
::f 0.6 a=0.0055
~ --Modeled k=1 kAim
0.4 o Experiment c=0.6
0.2 K,,=-4.5 kJ/m 3
Halong(OOI)
o
o 10 20 30 40 50
Appled magnetic filed (kA/m)

Figure 8. 16 Initial magnetization curves of iron.

case having an analytic function for the magnetization curve conveys significant
advantages in terms of computational time and complexity (Chiampi et al.,
1995) .
Comparisons between this model and the Preisach model are inevitable.
The former offers an analytical method for rapidly calculating the hysteresis
curves of materials. The latter offers a more complicated approach for fitting
almost any hysteresis curve to an arbitrary level of accuracy. Comparisons
have been made by Dupre et al. (Philips et aI., 1995; Dupre et aI., 1999;
Andrei and Stancu, 1999). Generally, it is found that computation in the Jiles-
Atherton model is faster than in the Preisach model. It is also found that by
appropriate choice of model parameters there is agreement between the two,
and that inevitably because of its greater degree of freedom, the Preisach is
able to produce magnetization curves that are in closer agreement with
experimental results.

8.7.5.2 Hard Magnetic Materials


Recent applications of the model to the NdFeB system have also been made.
This material forms the basis of an extremely important class of hard magnetic
materials. Compared to other permanent magnet materials, NdFeB has higher
Modeling of Hysteresis in Magnetic Materials 399

1.0

0.8 M s=505 kA/m


a=575 Nm
::t 0.6 a=0.0055
~ - - Modeled k=1 kNm
0.4 o Experiment c=0.6
K a=-4.5 kJ/m 3
0.2
Halong( 11 0)
o
o 10 20 30 40 50
Appled magnetic filed (kA/m)

l.0

0.8 M s=505 kNm


o a=575 Aim
.;{ 0.6 a=0.0055
~ - - Modeled k=1 kA/m
0.4 o Experiment c=0.6
K a=-4.5 kJ/m 3
0.2
Halong(OOI)
o
o 10 20 30 40 50
Appled magnetic filed (kNm)

Figure 8. 17 Initial magnetization curves of nickel.

coercivity and maximum energy product. Magnets based on NdFeB can be


manufactured from melt spun NdFeB, hot pressed and then fabricated into
polymer bonded magnets. The magnetic properties can be enhanced through
texturing by aligning the particles. This is done by preparing the magnetic
material in powder form through rapid-quenching techniques and then applying
a magnetic field while the polymer is curing.
Research has been performed by Chang and Shyu (1993) to study the
influence of packing factor on coercivity of particle arrays. However, it is still
unclear precisely how the compacting processes influence the magnetic
properties. Fang et al. (1998) have applied three different compacting
processes forming particles into. CD a free-movement state, (2) a randomly-
oriented state and Q) a field-oriented state. It has been shown that the sample
with particles in a free-movement state had the smallest coercivity of He = 160
kA/m while the other two had He = 400 kA/m. The main reason is that in the
free-movement state little external energy is needed to realign the particles
into the external field direction, while larger energies are necessary to
overcome the magnetocrystalline anisotropy energy in the randomly-oriented
and field-oriented particles. For the purposes of modeling, samples with
particles in the free-movement state and in the randomly-oriented state were
400 D.C. Jiles, X. Fang and W. Zhang

1.0

0.8
M s=142 kA/m
::;t 0.6 a=5 kAim
o Experiment
~ a=O.OOI
0.4 --Modeled k=5 kAim
c=O.1
0.2
K a= 470 kJ/m 3
Halong(OOI)
o
o 200 400 600 800
Appled magnetic filed (kAlm)

1.0
o 0
o Experiment o
0.8
--Modeled o
o M s=142 kA/m
:::{ 0.6 a=5 kA/m
~ a=O.OOI
0.4 k=5 kAim
c=O.1
0.2
K a=470 kJ/m 3
Halong(lIO)
o
200 400 600 800
Appled magnetic filed (kA/m)

Figure 8. 18 Initial magnetization curves of cobalt.

treated as isotropic materials, while the sample in the field-oriented state was
treated as an anisotropic structure with a degree of texture. The interactions
among the particles are long range, and therefore they were modeled using a
mean field method.
In the field-oriented specimens (F ig. 8. 19) the particles were al igned in
chains which led to a strong interactive coupl ing field and resulted in the
largest value of coupling coefficient a. For the randomly-oriented specimens
(Fig. 8.20), the particles were uniformly dispersed and isolated from each
other by the polymeric matrix. Because of the random distribution of the easy
axes, it was assumed that every single particle couples equally with its
neighboring particles. This resulted in a value of a that was only one third of
the value obtained in the case of the field-oriented material. For the free-
movement state particles, the coupling coefficient was found to be
intermediate between the two other cases. This is because the particles are in
contact with each other and can adjust their easy axis in three dimensions to
align with their neighbors, which caused a local alignment of magnetization in
small clusters.
NdFeB-based magnets have been limited by the low Curie temperature of
the hard magnetic phase RE 2 Fe14 B, which results in undesirable temperature
dependence of the coercive field as the Curie point is approached. Partial
Modeling of Hysteresis in Magnetic Materials 401

substitution of Dy for Nd and the addition of small amounts of other elements


reduced the effects of temperature and increased the reversible temperature
coefficient of the remanence B,. Mixing of TiO z , Dyz 0 3 or Ti high melting point
powders with REFeB and heat-treating at 500'C to 1200'C in an inert gas
atmosphere rei ieved internal distortion and caused surface modification,
leading to a markedly improved coercive force He. Fang et al. (1998) have
related these changes in composition and heat treatment to the magnetic
properties using the anisotropic extension of the model.
Gas atomization was performed on a close-coupled annular feed atomizer.
The samples were heated from room temperature at 2'C /min heating rate to
the heat treatment temperature, held for 600 s, and then furnace-cooled. Due
to the mass of the water-cooled furnace, the cool ing rate was relatively fast.
Modeled curves are shown in Figs. 8.21 and 8.22.

Figure 8. 19 SEM image of sample in field oriented state (Fang et aI., 1998).

Figure 8. 20 SEM image of sample in randomly oriented state (Fang et aI., 1998).
402 D.C. Jiles, X. Fang and W. Zhang

600
a=650 kAhn
k=700 kAhn
400 c=0.6

~
<C 200
[Isotropic]:
x=2.4
-""
~

~ [Anisotropic] :
c,
0
0 x=3.9
o Anisotropic
.~ K a=4.2 MJ/m 3
.!::! o Isotropic
1=0.44
v -200 Field axis: [100]
- Modeled
"
OJ)
oj

~
-400

-600 L----.l,-----_ _---L ...I.-_ _----.J ----L_


-2 -1 0 I 2
Magnetic field, H (MAim)

Figure 8. 21 Modeled and measured hysteresis curves of polymer bonded magnets


without heat treatment (Fang et aI., 1998).

a=1080kA/m
800 k=970kA/m
c=0.6
[Isotropic]:
~ 400 x=2.5
-""
~ [Anisotropic]:
~ x=1.4
0 1=0.7 o Anisotropic
.~ K a= 4.2 MJ/m 3 o Isotropic
.~
Field axis: [001] - Modeled
51
oj
-400
~

-800

-2 -1 0 1 2
Magnetic field, H (MAim)

Figure 8.22 Modeled and measured hysteresis curves of polymer bonded magnets after
heat treatment (Fang et aI., 1998).

Heat-treatment and the addition of Dy enhance the magnetic properties of


NdFeB materials mainly through the increase of the dissipation coefficient k
and the decrease of the reversibility c. The introduction of Dy into NdFeB
materials forms DY2 Fe14 B, which is a two-element, completely dilute system.
However, the lattice parameters change linearly with the Dy content which
resulted in different strengths of inter-atomic interactions and therefore
different coercivities were obtained. The heat treatment included quenching
and fast furnace cool ing which caused residual stress in the particles, thereby
Modeling of Hysteresis in Magnetic Materials 403

increasing coercivity and remanence. It was also shown from model


calculations that the domain density a increased after heat treatment, which
tends to reduce the coercivity and remanence, and that the coupling coefficient
a did not show any apparent dependence on the chemical additives or heat
treatment. This may be a result of its strong dependence on compacting
processes.
8.7.5.3 Modeling Stress Effects in Magnetostrictive Materials
Reliable model descriptions of magnetomechanical hysteresis materials et al.
have been developed by Calkins et al., Smith and Flatau (2000) which
provide simulations of transducer performance. The model was used to obtain
a calculated B, H hysteresis loop of the material, and then used a quadratic
relationship between magnetostriction and magnetization to determine
di\ / dM. In this way a hysteretic relationship between magnetostrictive strain
and applied magnetic field was obtained. The model was able to describe
strain under both major and minor loop excursions of magnetization versus
magnetic field.
Dapino et al. (2000) have shown how to include non-linear and hysteretic
effects directly into the transducer equation for strain. In this case, both the
elastic compliance and the piezomagnetic coefficient d were allowed to be
variable and path dependent. This generalization of the magnetostrictive
model is suitable for describing the behavior of magnetostrictive transducers at
high drive levels by providing a more complete description of the relationship
between input current to the coil and output strain of the transducer.

8.8 Summary

This paper has described the underlying basis for several hysteresis models
that can be used to describe the magnetic properties of materials. These
provide a diverse range of modeling capabilities that span length scales from
the discrete atomistic scale through nanoscopic and continuum/microscopic up
to the macroscopic everyday scale of devices and components. Examples of
modeling have been focused on specific cases of soft magnetic materials based
on iron, cobalt and nickel, and hard magnetic materials such as the NdFeB
system. These represent important classes of magnetic materials. Effects of
anisotropy, stress, frequency of excitation, compacting processing, chemical
composition and heat treatment have been related to these models.
Experimental results and simulation data have shown that these effects have
impacts on the magnetic properties and that these effects can be described
and understood through hysteresis model parameters.
404 D. C. Jiles, X. Fang and W. Zhang

References
Alessandro, B., C. Beatrice, G. Bertotti and A. Montorsi. J. Appl. Phys. 68:
2901 (1990)
Andrei, P., A. Stancu. J. MMM. 206: 160 (1999)
Antropov, V., B.N.Harmon and G.M.Stocks. Phys. Rev. 854: 1019 (1996)
Becker, R. Phys. Zeits. 33: 905 (1932)
Becker, M. Underlying theory of ferromagnetic hysteresis and coercivity.
Hirzel, Leipzig (1943)
Bertotti, G. IEEE Trans. Mag. 24: 621 (1988)
Bertotti, G. Energetic and thermodynamic aspects of hysteresis. Phys. Rev.
Letts. 76: 1739 (1996)
Bertotti, G. Hysteresis in Magnetism. Academic Press, San Diego (1998)
Bertotti , G., I. D. Mayergoyz, V. Basso and A. Magni. Functional integration
approach to hysteresis. Phys. Rev E 60: 1428 (1999)
Calkins, F. T., R. C. Smith and A. B. Flatau. IEEE Trans. Mag. 36: 429
(2000)
Chamberlin, R. V. Nature 408: 337 (2000)
Chang, Ching-Ray. Jyh-Pone Shyu. Particle interaction and coercivity for
acicular particles. Journal of Magnetism and Magnetic Materials 120 (3):
197 - 199( 1993)
Chantrell, R. W. Hysteresis particulate materials. In: G. C. Hadjipanayis ed.
Magnetic Hysteresis in Novel Magnetic Materials. Kluwer, Dordrecht
(1997 )
Chen, P. J. and S. T. Montgomery. Ferroelectrics 23 (3 - 4): 199 (1980)
Chiampi, M. , D. Chiarabaglio and M. Repetto. IEEE Trans. Mag. 31: 4306
(1995)
Comptes Rendus Acad. Seances 255, 1709, 1962
Dapino, M. J., R. C. Smith and A. B. Flatau. IEEE Trans. Mag. 36: 545
(2000)
Dupre, L. R. , G. Bertotti and J. A. A. Melkebeek. IEEE Trans. Mag. 34:
1168 (1998)
Dupre, L. R., O. Bottauscio, M. Chiampi, M. Repetto and J. A. A. Melkebeek.
IEEE Trans Mag. 35: 4171 (1999a)
Dupre, L. R., G. Ban, M. Von Rauch and J. A. A. Melkebeek. J. M. M. M. 195:
233 (1999b)
Dupre, L.R., R. Van Keer and J.A.A.Melkebeek. J.Appl.Phys. 85: 4376
( 1999c)
Escobar, M. A., R. Valenzuela and L. F. Magana. J. Appl. Phys. 54: 5935
(1983)
Escobar, M. A., L. F. Magana and R. Valenzuela. J. Appl. Phys. 57: 2142
(1985)
Fang, X., D. C. Jiles and Y. Shi. Journal of Magnetism and Magnetic Materials
Modeling of Hysteresis in Magnetic Materials 405

187: 79 (1998a)
Fang, X., Y. Shi and D. C. Jiles. IEEE, Transaction on Magnetics 34: 1291
(1998b)
Fidler ,J., T.Schrefl. J.Phys. D33: R135 (2000)
Gilbert, T.L. Phys. Rev. 100: 1243 (1955)
Globus, A. , P. Duplex. IEEE Trans. Mag. 2: 441 (1966)
Globus, A. , P.Duplex. Phys. Stat. Sol. 31: 765 (1969)
Globus, A., P.Duplex and M.Guyot. IEEE Trans. Mag. 7: 617 (1971)
Globus, A., M. Guyot. Phys. Stat. Sol. B 52.: 427 (1972)
Globus, A., M. Guyot. Phys. Stat. Sol. B 59: 447 (1973)
Globus, A. Universal hysteresis loop for soft ferromagnetic material. In:
Proc. Europ. Physical society, conference on soft magnetic material. a,
p. 233 (1975)
Jiles, D. C., J. B. Thoelke and M. K. Devine. IEEE Trans. Mag. 28: 27
(1992)
Jiles, D.C. IEEE Trans. Mag. 29: 3490 (1993)
Jiles, D.C. Journal of Applied Physics. Vo1.76, (no.l0, pt.1), 15 Nov. pp.
5849 - 5855 (1994)
Jiles, D. C., M. K. Devine. Journal of Magnetism and Magnetic Materials,
Vol. 140 - 144, pt. 3, (International Conference on Magnetism, ICM '94,
Warsaw, Poland, 22-26 Aug. 1994 Feb. pp. 1881-1882 (1995)
Jiles, D. C., A. Ramesh, Y. Shi and X. Fang. IEEE, Transactions on
Magnetics. 33: 3961 (1997)
Jiles, D. C. Czechoslovak Journal of Physics. 50: 893 ( 2000)
Jiles, D. C., S. J. Lee, J. Kenkel, K. L. Mellov. Appl. Phys. Lett. 77:
1029 (2000)
Jiles, D. C., D. L. Atherton. Journal of magnetism and magnetic materials.
61: 48-60 (1986)
Jiles, D. C. Modeling the magnetic properties of materials. Magnetics Society
Distinguished Lecture (1997 - 1998)
Kersten, M. Zeits. Angew. Phys. 8: 496 (1956)
Kronmuller, H., R. Fischer, R. Hertel and T. Leineweber. J. MMM. 175: 177
( 1997)
Landau, L. D. , E. M. Lifschitz. Phys. Z. Sowjetunion 8: 153 (1935)
Lee, E.W. ,J.E.L.Bishop. Proc. Phys. Soc. (Lond.) 89: 661 (1966)
Liorzou, F., B. Phelps and D. L. Atherton. IEEE Transactions on Magnetics.
36(2): 418 (2000)
Mayergoyz, I. D. Mathematical models of hysteresis. New York, Springer-
Verlag, New York ( 1991)
Neel, L., J. de Phys. et Radium 5: 18 (1944)
Neel, L., J. de Phys. et Radium 20: 215 (1959)
Phillips, D. A. ,L. R. Dupre and J. A. Melkebeek. IEEE Trans MAG. 31: 3551
(1995)
Pasquale, M. , V. Basso, G. Bertotti and D. C. Jiles. J. Appl. Phys. 83: 6497
406 D.C. Jiles, X. Fang and W. Zhang

(1998)
Preisach, F. Magnetic after-effects. Zeits. f. Physik. 94: 277 - 302 (1935)
Ramesh, A., D.C. JilesandJ. Roderick. IEEE Trans. Mag. 32: 4234 (1996)
Sablik, M. J., H. Kwun, G. L. Burkhardt, D. C. Jiles. J. Appl. Phys. 61:
3799 (1987)
Sablik, M. J., D. C. Jiles. J. Appl. Phys. 64: 5402 (1988)
Schlesinger, M. F. Physics in the noise. Nature 411: 641 (2001)
Slater, J.C. Phys. Rev. 49: 537 (1936)
Spratt, G. W. D., P. R. Bissell, R. W. Chantrell and E. P. Wohlfarth. J. MMM
75: 309 (1988)
Stoner, E.C. Phil. Mag. 15: 1080 (1933)
Stoner, E. C. , E. P. Wohlfarth. Phil. Trans. Roy. Soc., Vol. 240 A: 599-
642( 1948) and IEEE Trans. Mag. 27: 3475 (1991)
Tomas, I. private communication
Tuszynski, J. A., B. Mroz, H. Kiefte, M. J. Clouter. Ferroelectrics 77 ( 1) :
111-120 (1998)
Vicsek, T. A question of scale. Nature 411: 421 (2001)

Ames Laboratory is operated by Lowa State University for the US Department of Energy under
control unmber W-7405-ENG-82. This work was supported by the Materials Science Division
of the office of Basic Energy Science, USDOE
9 Coarse -graining and Hierarchical Simulation of
Magnetic Materials: the Fast Multipole Method

P. B. Visscher

9. 1 Introduction

The previous two chapters have discussed the micromagnetic simulation of


magnetic materials. One of the most difficult problems in this area is dealing
with the long range of the magnetic dipole interaction. Traditionally this
problem is deal with by a mean field approach, assuming that the effect of
distant dipoles is to produce a magnetic field locally that is proportional to the
total magnetization of the system. There is a huge amount of literature on
mean field approaches, but its accuracy is often such as to produce only
qualitative rather than quantitatively correct results. Recently another
approach has been developed to deal mathematically with the problem of long
ranged interactions, known as the Fast Multipole Method (FMM). The original
mathematical development of this method is rather complicated, but it has a
simple physical interpretation for magnetic systems.
The Fast Multipole Method was developed initially for gravitationally
interacting systems (galaxies or systems of galaxies) (Appel, 1985; Barnes
and Hut, 1986; Greengard and Rokhl in, 1987). If the interaction between
each of the N (N - 1) /2 pairs of particles in an N particle system has to be
calculated separately, the time required is proportional to N 2 (this is referred
to as an "N 2 algorithm"). The idea of the fast multi pole method is to group the
particles into successively larger regions, and compute the field of a source
region at points far from the source using a multi pole expansion rather than
explicitly adding all the constituent particles. It can be shown that this reduces
the time consumption from order N 2 to N (or N log N, depending on the details
of the implementation). The FMM has been extensively used for systems of
electrostatically interacting systems (Shimada, 1993; Wang and LeSar,
1994) but has only recently been appl ied to magnetic systems, in spite of the
fact that in some ways it is better suited for magnetic systems than for
gravitational or electrostatic ones. This is because it relies on the far field due
to a source region being efficiently describable in terms of its multipole
moments. This description is most efficient if the multi pole moments are small,
especially those of low order I, whose fields decay slowly with distance as
408 P. B. Visscher

1/ (1+2. In a magnetic system the zero order moment of a particle or grain


always vanishes, whereas in a gravitational system the zero order moment
(the total mass) can never vanish. In electrostatic systems, charges tend to
arrange themselves to minimize the zero moment, but it will not generally
vanish. In magnetic materials, not only does the zero order vanish but the
magnetization tends to arrange itself into flux-closure configurations (in which
the divergence of M vanishes) to minimize the pole density and, therefore, the
magnetostatic energy. In such a region not only the zero moment, but all
multipole moments, vanish. In the fast multi pole method we need to divide the
system hierarchically into regions; if we do this by dividing a flux-closure
system along flux lines, these regions will have zero pole density and therefore
all multipole moments will be zero. Of course in a real system the divergence
of the magnetization is not exactly zero, so the multi pole moments are not
exactly zero, but they will tend to be very small, so the multi pole expansion of
the field will converge very rapidly.
In spite of this efficiency advantage, the FMM has not been widely used in
micromagnetics, probably primarily because of its complexity, specifically
CD it requires arranging the system into a hierarchical tree, ~ it requires the
use of spherical harmonics, @ for each field point, it requires keeping track of
which source regions are treated at the coarsest level of the tree, which
becomes a complicated bookkeeping problem. At subsequent levels. Of these
complexities, only the first (hierarchy) is really unavoidable, but as we have
observed above, this may be tell ing us something about the physics of the
system, and studying it may not be a waste of time. We will describe in the
next section a recent formulation of the FMM that largely avoids complexities
~ and @~it uses simple polynomials in x, y, and z in place of spherical
harmonics, and the bookkeeping is done automatically by a recursive-function
implementation. It is hoped that such simplifications will lead to more use of
the FMM in magnetic applications.
After describing the fast multi pole method in its simplest form in
Sections 9.2 and 9.3, we will describe in Section 9.4 some of the history of
the FMM. In Section 9.5, we will discuss a few applications to micromagnetics.

9.2 The Fast Multipole Method: Simplest


Implementation

We will begin by describing the basic ideas of the fast multipole method. We
lump sources of magnetic field into source cells, one of which is shown
schematically on the right in Fig. 9. 1. The + and - signs represent charges, the
arrows represent dipole moments; in the specific case of a micromagnetic
simulation of a ferromagnetic material, the moments would be attached to
Coarse-graining and Hierarchical Simulation of Magnetic. . . 409

computational cells within this coarse-grained cell. We need to know the field
produced by these sources at each point of the field cell shown on the left.
Rather than calculate the field of each source object at every field point (an
order-N 2 problem), we calculate the multipole moments of the source cell
once, and use these to calculate the field at each field point, using Eq. (9. 6) .
Actually we can do even better~calculating the field from the multi pole
moments is quite slow. Rather than do it separately for each field point, we
can calculate the coefficients of a power series (Taylor) expansion of the
multipole field within the field cell (a polynomial in the coordinates x and y of
the axis system shown in Fig. 9.1). These Taylor coefficients can be added
up over all the source cells, so that all we have to do for each field point is to
evaluate the Taylor expansion. Effectively we are lumping the field points into
a field cell and calculating the field at many of them at once, just as we lumped
the source points.

Source cell

Field
ce/
y'L x
+

:------- +

Figure 9. 1 Sketch illustrating lumping of sources and Taylor expansion of fields.

Note that we have not described how we will decide how large the source
and field cells are. This is a critical question, because a multi pole expansion
becomes very slowly convergent when one gets close to the source cell.
Clearly our Taylor expansion can only describe sUfficiently distant source cells.
It is easy to show (Visscher and Apalkov, 2005) that the truncation error in a
multipole expansion including multi poles of order< I is of order ex I where ex is
the opening angle of the source as seen from the field, defined as (radius of
source) / (source-field distance). There is a similar expression for the
truncation error of the Taylor expansion, involving the opening angle of the
field as seen from the source. Fixing our desired error and the order I
determines a maximum allowed opening angle; only sources far enough to
appear smaller than this will produce smooth fields that can be accurately
computed from a Taylor expansion. Thus for any particular cell, such as the
one labeled" this" in Fig. 9. 2a, there are near sources whose fields are not
smooth in "this" cell, schematically represented as the interior of a circle, and
far sources (the exterior of the circle) whose fields are smooth; we will refer
to the total field of all of these far sources as the "smooth field" in "this" cell.
The reader familiar with C++ will note that our use of the word "this"
corresponds to its usage in C++: the cell is a C++ class, which has various
410 P. B. Visscher

data members (its center position, its size, its total magnetic moment, and
now its "smooth field," which is a set of Taylor coefficients). We hope that
our use of "this" to label the field cell under consideration will not prove too
confusing to readers unfamiliar with C++.
To see how the FMM works, suppose we know the smooth field in the
parent cell in Fig. 9. 2b. This includes the effects of all the cells outside the
larger (solid-line) circle. To find the smooth field of "this" cell, we must add
the effects of the sources in the crescent-shaped area between the circles,
which were not included in the smooth field of the parent. There is a technical
detail here-the known Taylor expansion of the smooth field in the parent cell
is with respect to the center of the parent, so we need to shift the origin of the
expansion to use it for the child "this." The reader will note that in any real
problem there is the smallest cell we wish to deal with (the computational cell
for which the magnetic moment is being computed, in the case of
micromagnetics). If "this" is one of these smallest cells, the FMM cannot
account for the near sources in the circle of Fig. 9. 2a by considering children of
" this." Thus these sources must be accounted for "by hand" ; in
micromagnetics this is done with a short-range micromagnetic kernel (see for
example the previous chapter). In an electrostatic calculation involving point
charges, this is done using Coulomb's law for the field of each near source. In
general the total number of such calculations is just of order N so this does not
increase the order of the overall algorithm.

Crescent
Far

Near .......

....... . .
......... .........--
~

(a) (b)

Figure 9. 2 (a) "this" cell and the circle representing its "near" sources; (b) the circle
for the parent cell. Sources in the crescent are far from "this" but near to the parent.

So far, our description of the fast multipole method applies to several


variants of the FMM. In the standard Greengard-Rokhlin implementation the
cells are in a rectangular lattice and the hierarchical subdivision subdivides
them by a factor of three in each direction, as indicated for the 20 case in
Fig. 9.3. This has two disadvantages, relative to the recursive implementation
we will describe. First, the level in the hierarchy must be kept track of "by
hand," whereas in a recursive implementation this is done automatically-the
information is kept in a function call stack, but the programmer does not have
to worry about it. Second, the maximum opening angle (and hence the
Coarse-graining and Hierarchical Simulation of Magnetic. . . 411

accuracy for a given multipole order) is fixed by the geometry at


2 tan- 1 (2 1/2 /3) , decreasing the flexibility of the method. Third, the
Greengard-Rokhlin form is only applicable to a system with a regular cubic
grid. Many micromagnetic systems are not exactly rectangular; for example
magnetic nanocomposites produced by ion implantation (Schulthess et al.,
2001) are randomly placed and difficult to describe by a regular lattice.

I.. ... ~ ~ ... .1.. . . " .. .:. .. 1: :..1


1 ; ,....1..1 .. .. ,..: j

Figure 9.3 The grid of the Greengard-Rokhlin FMM implementation. All the 3 x 3 coarse-
grained cells shown are "near" to the central grey one, everything outside them is "far".

Thus we will describe an alternative, recursive version of the FMM


(Visscher and Apalkov, 2001). We will assume a binary tree for simplicity,
although this is not essential. Thus the entire system is represented by the root
cell at the top of the tree in Fig. 9.4. The lowest cells are the computational
cells of a micromagnetic algorithm. The smooth fields in all the cells can be
calculated by a single call to a recursive function CalcFieldO which acts on (is
a member function of, for C++ devotees) the root cell. The smooth field of
the root cell is trivial, since there are no far sources (no part of the system is
far from the entire system.) This function calls itself recursively to calculate
the smooth field in the two children, etc. We will describe its functioning after
we have gotten down far enough that CalcField () is being called for a cell
(such as "this" in Fig. 9.2) whose parent has a nontrivial smooth field. We
shift the Taylor expansion describing this smooth field so it is centered on
"this" cell, and then add the fields due to the sources in the crescent region,
which are kept on a "partner list. " There is one partner Iist for each field cell.
This process continues recursively down the tree until the smooth fields in the
smallest cells have been calculated. In the smallest cells, we are not
interested in the complete Taylor expansion of the field, just the average
value, which is easily computed from the Taylor expansion. As mentioned
above, the contributions to the average field from the nearest small-cell
sources are calculated directly, by multiplying the source magnetization vector
by a stored matrix (the "kernel").
We should emphasize that the actual calculations carried out by the new
412 P. B. Visscher

Figure 9.4 Part of the binary tree describing a physical system for the FMM.

recursive algorithm are exactly the same as those of a more conventional


implementation using a binary or oct-tree (Brown et al., 2001); the advantage
of the recursive approach is only simplicity of programming-the CalcField()
function described above requires only a few lines. However, this is an
important factor, given that the perceived programming complexity of the FMM
has kept micromagnetic simulators using less efficient methods for many
years.
We have not, however, yet described how the partner list is generated.
This is again done recursively, by a function we will call CullPartnersO, which
is explicitly called only for the root cell. It starts with a list of all possible
partners, which for the root cell includes everything, i. e., it is the root cell
itself. We will call this list AIIPartners. For a general cell, AIIPartners is
passed to it by its parent and contains all sources not smooth over the parent-
this is the content of the larger circle in Fig. 9.2b. The desired partner list
contains only the subset of these that are smooth over "this" cell (the crescent
area). So the function Cull Partners () traverses the AIIPartners list, moving
anything satisfying this criterion into the partner list. What is left must be
passed to the two children of "this" cell, becoming their "AIIPartners" lists,
just before recursively calling CullPartnersO for the children. If we think about
initializing this recursion with the root cell, for which AIIPartners consists also
of the root cell, it is clear that this will never be far from any child, i. e., we
can never put it into a partner list. The solution to this problem is that if a cell
on the AIIPartners list is too close to be put on the Partner list, it might be just
because it is too large-one of its children might be far enough to produce a
smooth field. We don't want to make it much smaller than" this" cell, but if it
is larger we will replace it by its two children, and attempt to put them on the
Partner list.
The function CullPartnersO translates into very simple code. Although it
has taken us 20 lines of text to describe it above, it can be implemented in
about half that number of lines of code. The best way to understand it is to
apply it to a simple 2D system, as in Fig. 9.5.
In parts (a) - (c) of Fig. 9. 5, we are moving down the tree call ing
Coarse-graining and Hierarchical Simulation of Magnetic. . . 413

Partner Partner
This=Root

This
This This

(a) (b) (c) (d)

Part PaP
Partner
Part
This Partner
This rn lPalPa
Part
Partner ~~
"Part Part
(e) (f) (g)

Figure 9.5 Illustration of how CuIiPartners() generates the partner lists recursively as we
move down the tree. Source cells on the partner list of "this" cell are labeled" partner,"
"part," or "pa." The sources outside the heavy line are "far" sources.

Cull Partners () of successive descendants, and the AIlPartners list of near


sources (inside the heavy line) remains essentially the same, although its
member cells are being subdivided so the list gets longer. In part (d), finally
two of the cells on the parent's AIlPartners list (the cells at the top labeled
"partner") are far enough away to be put into the partner list of "this" cell. At
each later stage a few cells are similarly put on the partner list. The five
neighbors of the cell labeled "TH" at the last stage (g) cannot be treated by
the FMM without decreasing the computational cell size, and are handled by
hand using a stored kernel.
The important point here is that the interacting pairs of cells shown in
Fig. 9.5 are basically the same ones that would be used in a non-recursive
FMM, and not too different from those of the Greengard-Rokhl in
implementation. The difference is that the recursive algorithm generates them
automatically from about 10 lines of logically simple code, whereas in the
Greengard-Rokhlin approach the programmer has to generate a sketch like
Fig. 9.5 by hand and hard-wire the geometrical relationships into a much more
complicated and bug-prone code. Another factor that simplifies the code in
C++ is that we don't have to write functions for manipulating partner
lists-these can be implemented as Standard Template Library (STU lists.
Note also that while the function CullPartnersO is fairly efficient, this is not an
issue since the lists have to be generated only once.
In the next section, we will describe another recent improvement on the
FMM method, the replacement of spherical harmonics by cartesian functions.
414 P. B. Visscher

9. 3 Cartesian Formulation of the FMM

The fast multi pole method has usually been implemented in the past by
expanding the multipole fields in spherical harmonics. The apparent reason for
this is that the number of spherical harmonics is of order /2 (specifically,
(/ + 1)2), while the number of Cartesian monomials to the same order is of
order /3 (specifically, (/ + 1) ( / + 2) ( / + 3) /6) (Visscher and Apalkov,
2001). At first sight this appears to give the spherical harmonics an
overwhelming advantage as / becomes large. However, if we look at the
actual numbers (Table 9. 1) we see that to quadrupole order (which we might
want to use for rough estimates) the difference is only 10%. Even at 7th
order, which is as high as many fast-multi pole practitioners ever go, the
difference is not yet a factor of two. Since the coefficients of spherical
harmonics are complex, we can store a Cartesian multipole with a smaller
number of floating point variables than a spherical harmonic.
Table 9.1 The number of independent spherical harmonics up to order I, and the number of
monomials in the Cartesian coordinates to the same order.
0 2 3 4 5 6 7
Spherical 4 9 16 25 36 49 64
Cartesian 4 10 20 35 56 84 120

Balancing this small difference in the number of components against the


overwhelming simplicity of Cartesian monomials, the latter are certainly easier
to program and probably also more efficient. A Cartesian approach was
implemented to low order by hand several years ago (Shimada et al., 1994)
but this is, of course, awkward and bug-prone, so most implementations
continued to be based on spherical harmonics. Recently, however, a general
algorithm for computing convolution coefficients automatically in Cartesian
coordinates was developed (Visscher and Apalkov, 2001).
Because the spherical-harmonic implementation has been described many
times (e. g., Greengard and Rokhl in, 1987; Pfalzner and Gibbon, 1998) we
will describe the Cartesian implementation here. The overall structures of the
two methods are similar, in that both involve storing arrays of coefficients and
performing convolutions on them to calculate fields of multipoles or to shift the
origin of multipoles.
We will use the shorthand notation r n for an arbitrary Cartesian monomial
xnxynyznZ, where n = (n x ,n y ,n z ) is a vector of non-negative integers. Then

a Taylor expansion of the magnetic scalar potential V (r) can be written


Coarse-graining and Hierarchical Simulation of Magnetic... 415

VCr) = ~n --;vnr n
n.
C9.1)

where V n is the Taylor coefficient of V. Here we have defined the factorial of


a vector as the product of the factorials of its components, n! = nx! n y ! n z ! .
We will also use the shorthand notation n for the total order nx + n y + n z In
the FMM, we must store the Taylor expansion of the smooth field in each cell;
to order for 1=4, this is an array of 35 real numbers (from Table 9. 1) .
We also need to define Cartesian multipole moments, as

C9.2)

where pC r) is the magnetic pole density. In a magnetic system the zero order
moment 0(0.0.0) vanishes for any physical object, but we may want to consider
subcells with nonzero magnetic pole strength, so we will not exclude this term.
One of the operations we will need is a shift of origin: it is straightforward
to show that the multi pole moments 0' about another origin c are related to 0
by a convolution

"'" c,p0 , n-p'


On = LJ C9.3)
p p.
The corresponding shift in the spherical-harmonic formulation is also a
convolution, but it involving a sum over angular momentum indices I and m
(the angular momentum addition formulas familiar in quantum mechanics). The
present Cartesian formula is considerably simpler to implement.
The other required operation is the computation of the Taylor expansion
coefficients V n from the multipole moments Csometimes called the M2L, or
multipole-to-Iocal, conversion). This is easy to derive from Coulomb's law

VCC)=f pCr) d3 r C9.4)


I c - r I
by expanding 1/1 c - r I in a Taylor series whose coefficients are the
derivatives of 1/1 r I
C9.5)

These derivatives are polynomials that can easily be calculated by hand to low
orders CShimada, 1994) or by computer to arbitrary order CVisscher and
Apalkov, 2001). The resulting formula for the Taylor coefficients of the
potential near a point c due to a multi pole moment 0 at the origin is

Vm = C- 1) m ~D p+m Cc) 0 p . C9.6)


p

Again, the corresponding spherical-harmonic expression is also a convolution,


but a much more complicated one to compute. Although Eq. C9.6) may look
unfamiliar, it is just a generalization of the usual formulas of magnetostatics.
416 P. B. Visscher

For example, if Q is a dipole, the first order Taylor coefficient of V (i. e., the
magnetic field) has x -component
Hx =- V x = DxxQ x + DxyQ y + DxzQ z (9.7)
where 0 xx = (y2 + Z2 - 2x 2 ) / r 5 is the usual micromagnetic kernel. The two
convolution operations Eqs. ( 9. 3) and (9. 6) are all that we need to
implement the FMM.

9. 4 History of the FMM

The idea of lumping sources together into a larger source originated in


astrophysics, in the context of simulation of gravitationally interacting
galaxies. The first publication of this nature (Appel, 1985) was based on a
1981 undergraduate thesis, and involved lumping only into monopoles, not
general multi poles . However, it is worth noting because Appel's algorithm
allowed a general tree structure, not a rigid geometrical subdivision of space
such as in Fig. 9.4; for many years almost all subsequent work abandoned this
in favor of rigid geometries. Also, Appel's work was the first method to
reduce the time consumption from 0 (N 2 ) to 0 (N log N). A tree algorithm
based on a geometrical subdivision of space, that lumped more and more
distant gravitating sources into larger and larger cells with monopoles, soon
followed (Barnes and Hut, 1986), but these authors also did not lump field
cells, i. e., compute and shift Taylor expansions. The first complete fast
multipole algorithm was implemented in 20 (where the complication of
spherical harmonics is avoided by complex-variables techniques) (Greengard
and Rokhlin, 1987). It was subsequently shown (Esselink, 1992) that by
lumping the field as well as the source objects, the Greengard-Rokhl in method
is actually 0 (N) rather than 0 (N log N). The Greengard-Rokhlin approach
(with rigid geometry) was then implemented in 3D (Schmidt and Lee, 1991).
Schmidt and Lee computed the breakeven point (at which the FMM becomes
more efficient than direct summation over all pairs of particles) and found the
somewhat disheartening value of 70,000 particles. However, subsequent
authors have found lower breakeven points (several hundred particles or
cells), partly a result of demanding much lower accuracy (Schmidt and Lee
used multipole order 20). Also, the breakeven point is higher in point-particle
problems than in continuum problems such as micromagnetics (Seberino and
Bertram, 2001) because in a point-particle problem the FMM is competing
with a simple exact procedure (direct summation of pairs), which has no
direct analog in the continuum case, where deal ing with short-range
interactions properly is always a problem. Some of the relevant parameters
for these early calculations are shown in Table 9.2.
Coarse-graining and Hierarchical Simulation of Magnetic... 417

Table 9.2 Relevant parameters for early FMM and other source-lumping calculations
Authors Tree Order Angle CPU time Basis
2
(direct summation) Adaptive 0 0 N

Appel Cubic grid 0 -0.01 N log N

Barnes & Hut Cubic grid 0 -0.01 N log N

Greengard & Rokhlin Cubic grid 20 O(D N 20 complex


Schmidt & Lee Cubic grid 20 O(D N Sph. Harm.
Shimada et al. Cubic grid 4-5 O(D N Cartesian

There has been some work on accelerating the convolutions of the


spherical-harmonic FMM by doing a Fast Fourier Transform in the index space
(J and m) (Greengard and Rokhlin. 1988; Ell iott and Board. 1996). This may
be helpful for applications in which high accuracy is needed. but for
micromagnetic calculations (and in general for multi pole orders 1< 16) this
FFT has little advantage.
Since the late 1990s. there have been many appl ications of the FMM in
computational electromagnetics. for example the calculation of parasitic
capacitance of microstrip lines (Pan and Chew. 2000) and radar scattering
cross sections (Song et al.. 1998). The latter case requires a variant of the
. FMM that solves the Helmholtz equation rather than the Poisson equation. The
FMM of the Poisson equation can be considered the zero-frequency limit. The
slower dropoff with distance of the radiation fields makes this a somewhat
different problem. which we will not consider here.

9. 5 Micromagnetic Applications of the FMM

The first application of the Greengard-Rokhlin fast multipole method to


micromagnetics was made by Yuan and Bertram (1992) in 20. An earlier
paper (Blue and Scheinfein. 1991) used the hierarchical multi pole idea to
calculate fields in 20. but calculated each field point independently rather than
hierarchically. A 30 implementation was not published until 2001 (Seberino
and Bertram. 2001; Brown et al.. 2001). An important deterrent to the
widespread application of the FMM in micromagnetics is the extreme efficiency
of the Fast Fourier Transform (FFT). This allows calculating the potential by
Fourier transforming the pole density. computing the potential in k-space
analytically. and then Fourier transforming back to real space. Especially
because many years of effort have gone into optimizing the FFT. it has been
very difficult to compete with. even though in principle it is less efficient for
large N (N log N rather than N). Fortunately for the FMM. the FFT has an
important limitation: it requires an exactly regular grid. Although in principle
418 P. B. Visscher

any system can be represented on a regular grid, there are important features
of micromagnetic systems that are hard to model with a regular grid. For
example, small fluctuations in the shape of an interface are thought to have a
large influence on micromagnetics (the "Neel orange peel effect"). Systems in
which the magnetic material represents a small fraction of the total volume and
is distributed irregularly are difficult to deal with using the FFT and may be
best handled by the FMM. An example would be granular systems such as
those produced by ion implantation (Schulthess et al., 2001). Another
deterrent to the application of the FMM is the high degree of programming
complexity of traditional implementations; it is hoped that simplifications such
as the Cartesian formulation and the recursive-function implementation will
make it easier to apply.
As examples of the successful implementation of the FMM method in
micromagnetics, we present some results (Brown et a!., 2001) on a system of
isolated pillars, of the sort that have been synthesized through
electrochemically-assisted scanned-probe deposition (Wirth et al., 1999) or
by deposition in alumite (Sun et a!., 2000). A Greengard-Rokhl in type 3D
spherical-harmonic FMM calculation was done which showed that long pillars
switch by end nucleation, after which a domain wall moves along the pillar. A
series of snapshots of this motion is shown in Fig. 9.6.
1.0 1.05 1.10 1.15 1.20 ns

Figure 9. 6 Several snapshots of a magnetic column at different times (1.0 ns, etc.)
after the application of a reverse field near the coercivity. Initially the sample is
magnetized upward (light). Nucleation of downward magnetization (dark) occurs first at
the bottom, later at the top (Brown et al., 2001, reproduced with permission) .

The FMM algorithm is well adapted for domain-decomposition


parallelization, in which each processor would be assigned a subtree of cells
corresponding to a compact region of space; this minimizes the necessary
message-passing at interfaces between these regions. Both Brown et al. and
Seberino and Bertram (2001) parallelized the FMM algorithm. Figure 9. 7
Coarse-graining and Hierarchical Simulation of Magnetic. .. 419

shows a schematic parallelization, showing the message exchanges necessary


(curved arrows).

Processor # I Processor #2 Processor #3 Processor #4

IL-~::::::'J~:'.-o 0 0
/\ /\
0 0 0

Figure 9.7 Schematic of domain~decomposition parallelization, showing communication


between processors at domain boundaries (Seberino and Bertram, 2001, reproduced with
permission) .

Seberino and Bertram also compared the CPU time required for the FMM
and direct pairwise summation, with the results shown in Fig. 9. 8. It can be
seen that the parallel FMM becomes more efficient than pairwise summation
with only 500 cells. Even the serial FMM becomes more efficient at 1, 000
cells. Of course, this is a relevant comparison only if the problem is not
amenable to the FFT; if the FFT is possible the breakeven value of N would be
much larger.

14
,-.. 12 .... NoFMM
~ -- FMM serial
."
10 ..... FMM parallel
c: 8
0
.~
6
a'0"
0.
4
u
2

0 500 1000 1500 2000


Nwnber of elements, N

Figure 9. 8 Dependence of CPU time on number of computational cells, for direct


pairwise summation vs FMM (serial or parallel) (Seberino and Bertram, 2001,
reproduced with permission) .
420 P. B. Visscher

References
Appel, A. W. SIAM J. Computing 6, 85 (1985)
Barnes, J. and P. Hut. Nature 324: 446 (1986)
Blue, J. and M. Scheinfein. IEEE Trans. Magn. 27: 4778 (1991)
Brown, G., M. Novotny and P. Rikvold. Langevin simulation of thermally
activated magnetization reversal in nanoscale pillars. preprint, 2001
Elliott, W. D. andJ. A. Board, SIAMJ. Sci. Comput. 17: 398-415 (1996)
Esselink, K. Information Processing Let. 41: 141-147, (1992)
Greengard, L and V. J. Rokhlin. Compo Phys. 73: 325 - 348 (1987)
Pan, Y. C. and Chew, W. C. Microwave and Opt. Tech. Lett. 27: 13
(2000)
Pfalzner, S. and P. Gibbon. Many Body Tree Methods in Physics.
Cambridge University Press (1998)
Schmidt, K. E. and M. A. Lee. J. Stat. Phys. 63: 1223 - 1235 (1991)
Schulthess, T. C., M. Benakli, P. B. Visscher, K. D. Sorge, J. R.
Thompson. F. A. Modine, T. E. Haynes, L. A. Boatner, G. M. Stocks
and W. H. Butter. J. Appl. Phys. 89: 7594 (2001)
Seberino, C. and H. N. Bertram. IEEE Trans. Magn. 37: 1078 (2001)
Shimada, J., H. Kaneko and T. Takada. J. Comput. Chem. 15: 28 - 43
(1994)
Song, J. M., C. C. Lu, W. C. Chew and S. W. Lee. IEEE Ant. and Prop.
Mag. 40: 27 (1998)
Sun, M., G. Zangari and R. M. Metzger. IEEE Trans. Magn. 36: 3005-
3008 (2000)
Visscher, P. B. and D. Apalkov. Simple recursive Cartesian implementation
of Fast Mu/tipole method. Preprint available at http://bama . ua. edu/ "-
visscher/mumag (2001)
Wang, H. Y. and R. LeSar. J. Chem. Phys 104: 4173 (1994)
Wirth, S., M. Field, D. Awschalom and S. von Molnar. J. Appl. Phys. 85:
5249 (1999)
Yuan, S. and H.N. Bertram. IEEE Trans. Magn. 28: 2031 (1992)

The author wishes to acknowledge the support of the National Science Foundation, grant
DMR-MRSEC-0213985.
10 Numerical Simulation of Quasistatic and
Dynamic Remagnetization Processes with Special
Applications to Thin Films and Nanoparticles

D. V. Berkav, N. L. Gam

10.1 Basic Micromagnetic Concepts and Main Energy


Contributions

Theoretical micromagnetics as founded by W. F. Brown (Brown, 1963) is a


basically quite simple phenomenology which allows us to evaluate the total
magnetic free energy E tot of any ferromagnetic body if geometry, material
parameters and the magnetization configuration of this body are known. In its
" minimal" version, micromagnetics takes into account four energy
contributions-energy in the external field (Zeeman, energy) E ex! ' energy due
to the magnetocrystalline anisotropy E an , the exchange stiffness energy E exch
and the magnetodipolar interaction energy of the magnetic moments of the
ferromagnet, known as the stray (or demagnetizing) field energy E dem :
E to! = E ex! +E +an E exch +E dem (10. 1)
The inclusion of other energy terms-like surface anisotropy or magnetoelastic
energy-is possible (and in many cases even necessary), but we are not
going to consider them here.
To make this chapter self-contained we shall now write down integral
expressions for all energy terms listed above. Such expressions may be
derived on a very general basis (see, e. g., Landau and Lifshits, 1985;
Brown, 1963) using only very few assumptions concerning CD the symmetry of
the energy expression with respect to the time inversion, ~ the crystal
symmetry (for the anisotropy energy), @ the invariance properties of the
exchange interaction with respect to the space transformations (for the
exchange stiffness part) and @ the expression for the dipolar interaction
energy well known from the classical field theory (for the stray field term).
The results for the energy terms considered in Eq. (10. 1) are

E ext =- fHext(r) M(r) dV (10.2)


v
E~~ = - fK (r) [mer) nCr) J2 dV (10.3)
v
422 D. V. Berkov, N. L. Gorn

E eXCh = fA(r). [('vm x )2 + (\lm y )2 + (\lm z )2]. dV (10.4)


V

E dem =- ~ fM(r) Hdem(r) dV (10.5)


v
where H ext is external, H dem is demagnetizing field, and M is the magnitization
vector. The demagnetizing (stray) field H dem can be calculated as a
convolution of the magnetization distribution inside a ferromagnet with the
dipolar interaction kernel:

H ()
dem r
= f 3e r [e rM(r')] -M(r')dV'
I r _ r' I 3 (10.6)
V

where e r denotes the unit vector e r = (r- r')/ I r- r' I.


Several comments are in order before we can proceed:
( 1) The anisotropy energy term Eq. (10. 3) is given for the uniaxial
crystallographic anisotropy and in the simplest case when only the first
anisotropy constant is non-zero. The general ization to any anisotropy form
(cubic, mixed, etc.) and to any number of the relevant anisotropy constants is
straightforward and can be found in any standard textbook on ferromagnetic
materials (Chikazumi, 1997).
(2) The exchange energy term written as in Eq. (10.4) assumes that the
tensor of the exchange coefficients A ik can be reduced to a single scalar
quantity A. Hence Eq. (10.4) is valid, strictly speaking, for cubic crystals
only (Landau and Lifshits, 1985). However, it is widely used for crystals with
arbitrary symmetry due to the common belief that in a "normal" ferromagnet
Eq. (10.4) is a good approximation.
(3) The first three contributions to the total magnetic energy, E exl' E an
and Eexch' are local. This means that they can be expressed as integrals (over
the body volume) of the corresponding energy densities-eext(r), ean(r) and
eexch (r) -and that these densities depend on the magnetization (and, may be,
on its derivatives) at the point r only. This is not the case for the last term in
Eq. ( 10. 1), the stray field energy, because the stray field H dem (r) present in
the integrand of Eq. ( 10. 5) depends on the whole magnetization configuration
of the entire magnetic body (Eq. (10. 6)). This circumstance makes the
computation of this energy part and the stray field, as we shall see below, the
most difficult problem in quasistatic micromagnetics.

10.2 Discretization Methods: Simplicity and Speed


Versus Exact Shape Approximation

After the total energy of a ferromagnet has been expressed via Eqs. ( 10. 1) -
( 10. 6) as the functional of its magnetization configuration {M ( r) }, the
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 423

solution of the main static micromagnetic problem (to find the equilibrium
magnetization configuration under a given external condition) is conceptually
very simple: we have "only" to minimize the total energy with respect to
{M(r)}. The {M(r) }-configuration which delivers a minimum to our energy
functional is the equilibrium state we are looking for.
It is, however, quite obvious (just by inspection of the energy functional)
that the corresponding task can be solved analytically only for a very few
simple special cases, and this has been already done a long time ago (Brown,
1963a). Any practically interesting problem requires numerical minimization of
the system energy so that the first step to be accomplished is the discretization
of the ferromagnetic body under consideration.
Two main alternatives naturally arise: CD the translationally invariant grid
(preferably rectangular, but regular triangular or hexagonal dicretizations do
not lead to any serious complications either) and CZl the arbitrary tetrahedron
mesh. Below we shall briefly discuss pro and contra for both methods.

10.2.1 Regular (Translationally Invariant) Grids


The obvious advantage of this grid type is its simplicity~to discretize the
computation region, no special mesh generator is required. It is enough to
specify only the total structure size and the cell number in each direction;
discretization complete. The less obvious, but by far more important
advantage is that the grid obtained this way is translationally invariant. This
means that the discretization cell size is constant and the distance between the
two given cells depends on their indices i and j (we use the 10 notation to
simplify the formulae) only via the difference I i - j I. For this reason the
magnetodipolar interaction coefficients W ij between these cells (arising from
the discretization of the integrals Eq. ( 10. 5) and Eq. ( 10. 6 depend also only
on the difference between their indices Wij = W 1i - jl . This fact allows (at least
in principle, see the discussion below) to treat the sums containing these
coefficients as discrete convolution. The immediate and extremely welcome
consequence of this fact is the possibility to evaluate these sums using the fast
Fourier transformation (FFT) technique. FFT reduced the operation count for
the demagnetizing energy evaluation of a system of N finite elements
(discretization cells) from -- N 2 (when doing a direct summation in a real
space) down to -- N 10gN, making the treatment of really large-scale
configurations (up to several millions of cells on a personal computer)
possible.
However, any regular grid in general (and rectangular grid in particular)
has an obvious drawback: neither curved borders of a ferromagnetic body nor
its polycrystalline structure can be adequately approximated with such grids.
Such an approximation is necessary, e. g., when the explicit implementation of
the micromagnetic boundary conditions is required (e. g., oM/On-'- = 0 in the
424 D. V. Berkov, N. L. Gom

absence of a surface anisotropy, where n-'- is a unit vector normal to the body
surface). Another example is a system where one expects the crystallite grain
boundaries to play an important role in the remagnetization process, as it is
the case in the hard-soft magnetic nanocomposites and hard polycrystalline
magnetic materials.
Several attempts have been made to overcome the difficulty concerning
the representation of curved borders. These attempts reach from a simple
proportional decrease of magnetic moment magnitudes of cells cut by such
borders (Berkov et al., 2000) up to the sophisticated embedded curve
boundary (ECB)' method where the finite difference operators for the border
cells are modified explicitly (Parker et al., 2000). In the latter paper it was
shown that the adequate approximation of the stray and exchange fields on the
curved border could be achieved. However, in ECB significant modification of
the grid mesh on the element borders is required so that the authors of (Parker
et al., 2000) had to use the method based on the solution of the Poisson
equation for the magnetic potential (instead of the FFT technique) to find the
demagnetizing field.
To our knowledge, no systematic tests were carried out to study the
effects of imposing the rectangular mesh onto the polycrystalline structure of
real materials. Corresponding simulations of materials where the average
crystallite size remains approximately constant for the whole system seem to
provide consistent results (Berkov et ai., 2000), but for magnetic composites
where two phases may have very different grain sizes (Hadj ipanayis, 1999),
the straightforward usage of regular grids is surely not the best method to
account for these very different length scales.

10.2.2 Tetrahedron Mesh


An obvious alternative to the regular grid is an irregular tetrahedron mesh. Its
main advantages are CD the ability to approximate virtually any body shape
with high accuracy using a moderate number of finite elements and C?> the
ability to handle very different length scales using the adaptive mesh
refinement which can be also performed during the simulations.
The price to pay is the introduction of tetrahedra with largely varying sizes
and shapes so that one has to refuse the FFT-based techniques to evaluate the
stray field Helem' As it was mentioned above, the evaluation of this field using
the direct summation of contributions from all other finite elements on the given
tetrahedron is an '" N 2 operation. Hence, such a summation is out of question
even taking into account that the number of finite elements N required by the
tetrahedron discretization may be considerably smaller than for the regular
mesh.
The most elaborate method to cope with this difficulty when solving
quasistatic problems is based on the idea not to evaluate the stray field at all.
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 425

There exists an alternative way to calculate the stray field energy based on the
variational formalism dating all the way back to Brown (1963) and revived
later by Asselin and Thiele (1986). Namely, it is straightforward to
demonstrate that the minimal value of the stray field energy E OOm is equal to the
minimum of the functional

W(A,M) = ~f [\7 x A(r) - 4lTM(r)]2dV <10.7)


Q

where the vector potential A is related to the stray induction as Boom = rot A.
Hence E oom can be found by minimizing this functional with respect to the vector
fields A(r) and M(r). The great advantage of this formulation is that both A
and M are local quantities, so that the total system energy does not depend
anymore on non-local fields like H oom ; the energy functional is entirely local.
Unfortunately, this formal ism also has several serious drawbacks. The
first (relatively minor) problem is that the number of independent variables has
been increased, because there are now two independent vector fields- A (r)
and M(r)-to deal with. A more serious problem is that the integration in
Eq. (10. 1) has to be performed over the whole space Q which is obviously
impossible to do numerically. Hence it is necessary to introduce an outer
border of the integration area and to map the rest of the universe into a
subdomain inside this border using some conformal transformation. The
position of the outer border and the optimal transformation must be determined
separately for each new geometry to be simulated. In addition, the region
between the magnetic body itself and the outer border must also be
discretized, thus increasing the number of finite elements. The technique
briefly outlined here was successfully applied to the study of the reversal
behavior of hard magnetic particles (Schrefl et aI., 1994a), magnetic
properties of composite magnets (Fischer et al., 1995; Schrefl et al.,
1994b), magnetic multi layers (Schrefl et aI., 1996) etc.
However, in many important applications-e. g., when studying the
magnetization dynamics-the stray field simply must be calculated, because it
is required to solve the corresponding equations of motion. In such cases the
most common method to compute the stray field H oom is to take the (numerical)
derivative of the scalar magnetic potential If>. It must be found from the
numerical solution of the corresponding Poisson equation
<10.8)
where the density of "magnetic charges" pmag(r) is defined as Pmag(r) = -div M(r).
Leaving aside the substantial but technical difficulty of solving such equations
on an arbitrary mesh and the problem of getting an accurate numerical
derivative of If> (r) on such a mesh, one is sti II confronted with the problem of
the same nature as by minimizing (Eq. ( 10. 7 that boundary conditions for
the potential If> are set on infinity: If> (r - 00 ) - O. Basically there exist two
techniques to avoid the usage of these conditions: CD the hybrid finite/
426 D. V. Berkov, N. L. Gom

boundary element method (FEM/BEM) and ~ the usage of the asymptotic


boundary conditions (ABC).
In the FEM/BEM approach (Fredkin and Koehler, 1990) the scalar
magnetic potential is split into two parts: 4> (r) = 4>, (r) + 4>2 (r). The first part
4>, Cr) is defined as being zero outside the ferromagnetic body and obeys the
Poisson Eq. (10.8) inside it. The boundary condition for 4> 1(r) is defined on
the ferromagnet's surface so that the problem of evaluating this part of
magnetic potential deals with the finite region occupied by the ferromagnet.
The properties of the second part 4>2 (r) follow from the definition of 4> 1(r) :
the function 4>2 (r) satisfies the Laplace equation in the whole space and
exhibits a jump on the ferromagnets surface (with the jump magnitude
governed by the 4>, (r) -values on this surface), thus representing a double-
layer potential which can be found by the corresponding surface integration.
The matrices arising by the numerical solution of the Poisson equation for
4>, (r) are sparse, which enables an application of appropriate highly efficient
techniques for the solution of the sparse algebraic systems. Unfortunately, the
matrices establishing the connection between 4>1 (r) and 4>2 (r) are full, and
their size is of the order N 8 X N 8 (where N 8 is the number of mesh nodes on
the surface of a ferromagnetic body). This makes the computation rather time-
consuming especially for thin films and nanodots, where the fraction of nodes
on the surface approaches 1. O. However, for moderate mesh sizes it was
possible to compute not only quasistatic magnetic structures of 3D particles
(see, e. g., Ref. (Koehler and Fredkin, 1992, but also dynamical
remagnetization processes of thin magnetic platelets (Schrefl et aI., 1997).
The asymptotic boundary condition (ABC) method uses the fact that
outside the ferromagnetic body the Poisson equation, reduces to the Laplace
equation, and thus its solution can be represented as a sum of, e. g., spherical
harmonics. This can be used setting the artificial boundary conditions for the
potential ()4>(r out )/()n-L = R 4>(r out ) set on the artificial boundary rout outside
the ferromagnet. The matrix elements of the operator R (which establ ish the
relation between the potential and its normal derivatives on rout) can be
calculated using spherical harmonics as a set of basis functions. For the exact
representation of 4> and R one needs the infinite number of them, but to
approximate the true solution only a finite subset of these harmonics is
needed. The number of spherical harmonics L sh required to achieve the
prescribed accuracy quickly decreases with increasing distance between the
body and the point where the solution must be found. The consequence is that
if rout is chosen far enough from the ferromagnetic body to be simulated, one
can escape with moderate L sh values. In this case the size L sh x L sh of matrices
describing the boundary conditions is much less then the number of boundary
elements (as it was for the FEM/BEM described above). The full algorithm
resulting from this idea is fairly compl icated and time consuming (see Ref.
(Yang and Fredkin, 1998) for detailed discussion of the corresponding
difficulties) and still too new to judge its real capacity (almost all large-scale
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 427

results reported in (Yang and Fredkin, 1998) were obtained on a


supercomputer Cray T3E) .
The interested reader will find a more detailed discussion of topics dealing
with the tetrahedron discretization in contributions to this volume written by T.
Schrefl and (up to some extenD P. Visscher. We have provided here a short
comparison of the two calculation techniques (regular grid and tetrahedron
mesh) because the choice of the discretization plays a crucial role in the
subsequent choice of the simulation algorithm. For this reason one should be
aware of all the advantages, limitations and consequences when choosing the
discretization method. Another reason for drawing the attention of the reader
to this problem is the absence (to our knowledge) of the systematic research
concerning the question of which discretization is best suited for which
magnetic systems, leaving apart some simplest obvious cases (like Bloch
walls in homogeneous thin films or composite nanomagnets with hard grains
and soft magnetic phase having very different characteristic length scales).
Such a systematic comparison of these two discretization techniques is clearly
necessary to ensure that the best available method will be applied in each
specific case.

10.3 Evaluation of Various Energy Contributions

In this chapter we discuss the evaluation of energy contribution Eqs. (10.2)-


( 10. 5) after the continuous problem of minimizing the micromagnetic energy
Eq. (10. 1) has been discretized, i. e., we discuss the finite difference
versions of the integrals Eqs. (10. 2) - ( 10. 6). Below we use the simplest
finite element approximation of the magnetization field, assuming that the
magnetization inside each discretization cell is constant. This constant vector
for the i-th cell is denoted as M; .
The discretization of the ferromagnet energy in an external field Eq. (10.2) is
straightforward: the corresponding finite-difference expression is

E ext =- ~ <H1xt > M;/j. V; (10.9)

where V; is the volume of the corresponding cell and <H1 xt > denotes the spatial
average of the external field (which can be non-homogeneous) over this cell.
Eq. (10.9) is obviously valid for any discretization method. Improvements of
this simplest approximation may be achieved taking into account the variation
of the magnetization inside the discretization cell; this question, being quite
simple and purely technical, will not be discussed here.
428 D. V. Berkov, N. L. Gom

10.3.1 Anisotropy Energy in Polycrystalline Samples


The discretized version of the anisotropy energy is also simple, because the
continuous form for the corresponding energy density e an depends-as the
energy density in an external field eeXI-on the magnetization values only (see
the integrand in Eq. ( 10. 3); this statement is true for any anisotropy type).
So instead of Eq. ( 10. 3) we have in the discrete version

(10. 10)

The indices i by the anisotropy constant K i and the anisotropy axis unit vector
nj remind us that both the anisotropy magnitude and the anisotropy axes
orientation may be cell-dependent.
The standard situation where such a dependence takes place is the
simulation of a polycrystalline material with the anisotropy varying (both in
magnitude and direction) from one crystallite to another. In this case it is
necessary to generate the corresponding polycrystalline structure within the
simulation volume to ensure that the influence of this structure will be taken into
account properly.
For this purpose we use the following procedure (Berkov et aI., 2000).
From the average crystallite size <0 > we evaluate the number of crystallites L
which should be placed in the simulation volume VIOl as L = V lot / <0 >3 and
place L growth centers randomly inside this volume (Fig. 10. 1a). Afterwards
we simulate the isotropic crystallite growth starting simultaneously from all
these centers (Fig. 10. 1b) and terminate the growth where and when two
crystallites meet each other (Fig. 10. 1c). The procedure is complete when
the whole space is covered with the crystallites (Fig. 10. 1d). To simulate this
process we use a much finer lattice than for the micromagnetic simulation
itself. After completion of the growth process we assign to each cell of the
"micromagnetic" lattice the anisotropy parameters (K j and nj) of that
crystallite which occupies the largest fraction of this cell.

(a) (b) (e) (d)

Figure 10. 1 Crystal growth procedure used in our micromagnetic simulations (20
demonstration, see text for details) .
When applied in this primitive form, the procedure described above is
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 429

equivalent to the standard Voronoi-Delaunay construction of the polyhedron


polycrystallites used, e. g., in some commercial micromagnetic packages
(ARM, MagFEM3D). However, when any additional information concerning
the crystallite growth is available (e. g., the anisotropy of the growth
kinetics), this information can be incorporated into our generation procedure
more easily than into standard polyhedra-generating programs.

10.3. 2 Exchange Energy: Node-supported Discretization,


Heisenberg-like Form and Angle-Based Interpolation
The finite-difference representation of the exchange term Eq. ( 10. 4) is more
tricky than the approximation of the external field and anisotropy contributions
because the exchange energy depends on the spatial derivatives of the
magnetization M (r). Hence the calculation of the exchange field in a given
lattice cell involves also the moments at the neighboring cells. Below we
consider the three most common ways to calculate the exchange energy on a
regular lattice.
10.3.2. 1 Approximation of the Gradients of the Cartesian Moment
Projections
The most straightforward method to discretize the exchange energy functional
is to use some sort of a standard finite difference approximation for the
gradient operators present in Eq. (10. 4). To approximate the exchange
energy for a lattice cell marked grey in Fig. 10. 2, we need the finite
difference expression for the average values of () m x / () x, () m x / () y, etc. inside
this cell. The simplest formula uses the linear interpolation between the
corresponding nodes and leads to the result

(10.11)

where AX denotes the lattice cell size in the x-direction. Similar formulae hold
for derivatives of other moment projections over other coordinates necessary
to evaluate the gradients in Eq. (10. 4). The accuracy of such a finite
difference representation for the exchange energy may be improved by using
the higher-order numerical approximations (Abramovitz and Stegun, 1968) of
the moment derivatives (which would obviously require the inclusion of the
moment projections of next nearest neighbors, etc.) Mostly the five-point
formula is used, for it represents the optimal compromise between the
computational time and accuracy (Labrune and Miltat, 1990; Berkov et aI.,
1993: Wright et al., 1997, etc.). The whole method is conceptually simple,
easy to implement and enables a fast evaluation of the exchange field H exch =
- 5Eexch /5M.
However, the approximation of the exchange energy using Eq. ( 10. 11)
430 D. V. Berkev, N. L. Gem

Figure 10. 2 A" normal" moment configuration suitable for the exchange energy
evaluation within the approximation Eq. ( 10. 10) .

and similar formulae also has serious drawbacks. The first disadvantage is that
by using a linear (or any other polynomial interpolation) for the Cartesian
coordinates between the nodes we violate the condition I M I = Const which is
one of the basic conditions in the physics of a ferromagnetic states when we
are far below the Curie temperature of the material (Landau and Lifshitz,
1985). Even when we overlook this unpleasant violation of one of the basic
principles, we are still left with the problem that such a violation unavoidably
leads to a systematic error by the exchange energy evaluation. This error
obviously decreases when the angle between the neighboring moments
decreases. However, to keep all these angles really small we must have a
very fine discretization lattice, and hence a large number of finite elements.
Another problem arising occasionally under certain unfavorable conditions
is much more serious. Namely, performing simulations of soft (low anisotropy)
magnetic materials in small external fields using the approximation Eq. (10. 11)
for E exch ' we have observed that the system tends to find itself in the so-called
checkerboard state (shown in Fig. 10. 3), in which the moments on the
opposite cell corners are aligned in an anti parallel way. The reason for this
alignment is pretty clear: in this state (m2 - m1) = - (m4 - m3) so that the
average derivative of any m-projection over any Cartesian coordinate
evaluated according to Eq. (10. 11) is exactly zero. For this reason the overall
exchange energy of the system is also zero, so that the checkerboard state
undoubtedly delivers the global minimum to the non-negative exchange energy
Eq. ( 10. 4). This result remains valid for any higher-order polynomial
approximations of the average derivatives of Cartesian moment projections.
And it comes even better: if for the evaluation of the stray field energy
one uses the "charge formal ism," where (fictitious) magnetic charges are
evaluated according to their definitions as pmag = - div M, the same state also
has a zero stray field energy (no charges!), i. e., the smallest possible value
of E dem' So, when a system to be simulated has a low anisotropy and the
external field is also low (e. g., we are interested in a remanent state of a soft
magnetic element), the absurd configuration in Fig. 10.3 is surely the most
energetically favorable state of the system-from "the point of view" of this
finite-element version.
The mathematical reason for this disaster is evident: the finite difference
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 431

Figure 10.3 The checkerboard moment configuration delivering the global minimum finite-
difference version Eq. (10. 10) of the exchange energy.

approximation Eq. (10. 11) for the function derivative holds (by its definition)
only for slowly varying functions, i. e., for functions changing their values only
slightly (compared to the values themselves) between the discretization
nodes. This is obviously not the case for the behavior shown in Fig. 10. 3,
which means that the finite difference approximation used here for the gradient
operator is absolutely unacceptab Ie. However, the computer program does
not know this and drives the system towards the checkerboard state whenever
it is able to overcome the (not very high) energy barrier separating the region
of physically reasonable states (where the angles between the adjacent
moments are small) from the checkerboard order. The only remedy is the
increase of the cell number and the limitation of the iteration step length by the
energy minimization ~both resulting in the substantial increase of the
computation time.

10.3.2.2 Usage of the Heisenberg Form of the Exchange Interaction

Another frequently employed form for computing the exchange stiffness energy
of the system (ARM, Zhu and Bertram, 1988; Berkov and Gom, 1998;
Bertram and Seberino, 1999, etc.) is the scalar product of the neighboring
magnetic moments

E exch =- ~ J ij (m i mj ) (10. 12)


(i ,j)

where the sum of the nearest neighbors < i, j > is taken. The exchange
interaction coefficients J ij may be site-dependent (e. g., to account for the
weakening of the exchange interaction on the crystallite borders) and are
positive for a ferromagnet.
Before we proceed with the analysis of this approximation, we point out
that Eq. (10. 12) is not, as it is often claimed, the "only correct generic form
for the exchange interaction in micromagnetics introduced by Heisenberg." It
is certainly true that the expression E exch = - J (Sl S2) was introduced by
Heisenberg in 1929, not to assist micromagnetics (which did not even exist by
that time), but just as the simplest form of the quantum mechanical exchange
interaction which obeys certain rules for the spin operators. What we need to
find out, whether Eq. (10. 12) is a good approximation for the continuous form
432 D. V. Berkov, N. L. Gorn

of the exchange stiffness energy Eq. (10. 4), and if yes, under which
conditions.
From this point of view, Eq. (10. 12) seems to be suitable: it can be
easily shown that for small angles between adjacent cell moments the finite-
difference approximation (Eq. ( 10. 12 on a rectangular lattice is equivalent
to its continuous counterpart Eq. ( 10. 4) if J xx =2 AVcell / I:ix 2 , where Vcell is the
lattice cell volume (similarly for Jyyand Jzz-coefficients). In addition, using
Eq. ( 10. 12) we do not have to worry about the violation of the condition I M I =
Const (or I m I = 1), as it was the case for the Eexch-representation based on
Eq. ( 10. 11): since we use the vectors mi on lattice nodes only and no
interpolation is necessary, this condition is always fulfilled. The third
advantage of the scalar product Eq. (10. 12) is even more important: the
exchange stiffness energy evaluated this way increases monotonously with the
angle between the adjacent moments, so that states like the ones shown in
Fig. 10. 3 do not correspond anymore to an energy minimum. The last
attraction of Eq. (10. 12) is its simplicity: the evaluation of the scalar product
is a fast operation and its derivatives (necessary to calculate the exchange
contribution to the effective field) are linear in moment projections.
However, Eq. (10. 12) also has an important drawback: it fails
quantitatively to approximate the exchange stiffness energy for the
configuration with large angles between adjacent moments. To demonstrate
this, it is sufficient to calculate Eexch of the simple magnetization configuration
shown in Fig. 10.4, where m (r) on the left and right sides Iie in the 0yz-plane
and are fixed. Let us assume that the magnetization inside the rectangular cell
rotates remaining in the Oyz-plane, and the rotation angle <P(x) (angle with
the Oz-axis) varies linearly along the Ox-axis: m x = 0, my = sin <P = sin (ax) ,
m z = cos <P = cos (ax).

z
y
O~-x--------I'
Ix

Figure 10.4 A simple magnetization configuration used to analyze the behavior of the
Heisenberg approximation Eq. (10. 14) for the Eexch' The magnetization rotates in the
Oyz-plane.

In this case the exchange integral Eq. (10. 4) can be evaluated


analytically with the result

(10.13)

where I:i V denotes the cell volume, I x -the cell size in the x -direction and
<Pc -the angle between the initial (x = 0) and final (x = I x) orientations of m.
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 433

The exchange energy for the same configuration evaluated using the
approximation (Eq. (10.12 with the properly adjusted coefficient J is
_ 2AI1V ,j"
E Heis
exch - -/~-cos't'c. (10.14)

(Obviously, it does not depend on the assumption how the magnetization


varies between the initial and final orientations. )
Comparison of these two results (using Eq. (10. 14) shifted to give E exch =
o for 4>c = 0) is shown in Fig. 10.5. Up to the angles'" IT/4 = 45 the
approximation Eq. (10. 12) performs really well (this was actually the purpose
of adjusting J). However, for larger angles the exchange energy given by the
scalar product is getting increasingly lower than the actual value, and for the
maximal angle 4>c = IT the correct value is almost 2. 5 times larger than the
approximated one.
10
8

o 2 3
cf> (rad)

Figure 10.5 Comparison of the Heisenberg approximation Eq. (10.14) for the exchange
energy with the correct results Eq. (10. 13) for the magnetization configuration shown in
Fig. 10.4.

One may object that this is only a demonstration of a well-known fact that
the second order Taylor expansion of almost any function works well only in the
vicinity of the expansion point and this trivial statement is not worth placing an
extra figure in a "scientific" contribution. Another argument against such a
comparison is that the finite element method under discussion works well
anyway only for configurations where all angles between adjacent moments are
small.
Both objections by themselves are correct. Indeed, the Taylor expansion
is valid near the expansion point only, but the computer program does not
know this applying Eq. ( 10. 12) for any configuration. It is also true that the
micromagnetic code works properly for magnetization configurations with small
angles between the adjacent moments (at least in most cases), but the check
that all these angles are really small is absent at least in many commercial
packages. And apart from these quite trivial facts there also exist more
important reasons to take seriously the underestimation of E exch by the (m!
m2 ) -form for large angles between the moments.
The first reason is that in many important cases the exchange interaction
434 D. V. Berkov. N. L. Gorn

between the neighboring cells is weakened. An important example of such a


situation is a polycrystalline sample with disturbed intergrain borders. In this
case we could safely allow larger angles between adjacent moments belonging
to different crystallites if we would possess an approximation describing such a
situation more adequately.
The second reason is that the dipole moment of the configuration with
adjacent moments nearly opposite to each other is almost zero-leading to a
very small stray field energy. This means that the algorithm based on Eq. ( 10. 12)
would be tempted to align the adjacent moments nearly antiparallel. because
this way it could greatly decrease the stray field energy at the cost of a
moderate increase of the exchange energy. We have indeed frequently
observed such a situation especially when modeling domain walls. Of course.
one would recover the correct solution decreasing the cell size, but it
obviously means increasing the cell number. And the third problem with the
scalar product form arises for the configuration with large angles ('" TT). If
such a state is created once during the simulations, it evolves very slowly,
because the effective field-being the derivative of the energy over the
moment orientations-is small for such large angles (see Fig. 10.5).

10.3.2.3 Approximation Using the Angles Between Adjacent Moments


Taking into account all the reasons listed above, we have adopted the
approximation of the exchange energy which uses the angles ex ij (m i , mj )
between adjacent moments:

E:~~~ = ~Ji/Kijlt..Vi't..Vj)ex7j' <10.15)


<i. j)

Here the exchange coefficient J ij in, e. g., x -direction, depends on the


exchange weakening K ij (caused, e. g., by the grain boundaries) and the
volumes t.. V i and t.. V j of the adjacent rectangular cells as J ij = AK ij (t.. Vi +
t.. V j) / (2t..x 2 ). The factor (t.. Vi + t.. V j ) /2 appears because in our
discretization scheme the moment vectors are located in the middle of the
corresponding cell, so that the volume between the moments of cell i and j
(compare with Fig. 10.4) is (t..V i +t..V j )/2.
The only drawback of this approximation is a somewhat complicated
evaluation of the exchange field MXCh = - () Eexch/d m i (the computation of the
acos ( ) or asin ( ) derivatives is necessary, depending on the method
applied for the evaluation of the angle ex ij (mi' mj) from the moment
projections), but this turns out to be only a minor correction to the total
computation time.

10.3.3 Stray Field Evaluation on Regular Grids


As it was mentioned earlier, the computation of the magnetodipolar interaction
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 435

energy and the corresponding field is the most difficult problem in numerical
micromagnetics due to the long-range nature of this interaction. Fortunately,
for the discretization using regular lattice this computation can be performed
with methods based on the fast fourier transformation (FFT) techniques both
for systems with open boundaries (Yuan and Bertram 1992; Berkov et aI.,
1993) and periodic boundary conditions (Mansuripur and Giles, 1989; Be'rkov
and Gorn, 1998). The implementation in the latter case is more sophisticated
and requires the combination of the FFT technique with the Ewald method
(Berkov and Gorn, 1998). Below we shall briefly discuss both kinds of
systems.
10.3.3.1 A Finite Systems (Open Boundaries): FFT-Methods with
Zero-Padding
The usage of the FFT-based methods for the stray field evaluation is based on
the observation that, e. g., for a lattice of point dipoles JJ the field on the i -th
lattice site created by all other dipoles
3ei/eij JJj) - JJj
H; = (10.16)
t,rIj

can be rewritten as

H "I = ' " W"./3


L..J I1 J3
I/""J '
(10.17)
/3.i'cfti
Here cx,/3=x, y, z, unit vectors e ij in Eq. (10.16) are defined as eij =t,rij/
t,r ij and vectors t,rij = ri - rj connect the lattice sites i and j (again, we use
10 notation for simplicity). Analogous formulae can be derived also for
rectangular cells of finite size under assumption that the magnetization inside
each cell is homogeneous even for more complicated magnetization behavior
(corresponding expressions for the interaction matrices Wij can be found in
(Ramst6ck et al., 1994 .
It is evident, that for the translationally invariant (regular) lattice the
elements of the interaction matrices W ij depend on the site (cell) numbers i
and j via their difference I i - j I only. In this case Eq. (10. 17) for the dipolar
field components

H "i -- ' " W"/3


L..J I ;- jI JJ /3
j (10.18)
/3.i'cft;

can be recognized as a discrete convolution: in Eq. (10. 18) the set of the
magnetization components {JJf} plays the role of a signal and the interaction
matrices W jl represent a response function in terms of the signal processing
1
;-

where the whole formalism comes from. Such a convolution is normally


performed via the FFT using the well-known theorem that under certain
conditions (see below) the Fourier transform H" (k) of the stray field is the
dot product of the Fourier transforms of the interaction coefficients W"/3 (k)
and magnetic moments JJ/3 (k) :
436 D. V. Berkov, N. L. Gorn

H"(k) = ~W"~(k). 1J~(k). (10. 19)


~

The gain is evident: Fourier components of the magnetization configuration can


be evaluated all simultaneously in'" N log N operations (N being the total
number of lattice cells) using the FFT algorithm. The inverse Fourier transform
from H" (k) to the field components in real space H" (r) requires the same
effort. Finally, the multiplication Eq. ( 10. 19) requires only'" N operations,
because for N lattice sites we have N independent Fourier components. Hence
the evaluation of the dipolar field for all lattice cells can be performed in '"
N log N operations only (instead of N 2 for a direct summation via Eq. ( 10. 17)).
This means a nearly linear dependence of the calculation time on the cell
number-almost as in systems with the short-range interaction. Using this
technique large scale 2D and 3D micromagnetic simulations (with the cell
number N '" 105 - 106 ) have been performed (Yuan and Bertram, 1992;
Berkov et aI., 1993; Wright et aI., 1997; Hubert and Rave, 1999, etc. ) .
The implementation of this basically very simple idea is not as
straightforward as it seems to be due to the conditions implied by the
convolution theorem. The first condition (in terms of our calculations) requires
that the magnetization configuration should be a periodic function (in space).
The second condition requires that the interaction function should have a finite
range, could be at least cut off at some finite distance. Both conditions are
violated in most micromagnetic applications. First, the magnetization
configuration is, generally speaking, not periodic and secondly, the dipolar
interaction is a long-range one: due to its quite slow decay ('" r - 3) no
physically reasonable cut-off radius can be introduced.
The conceptually simplest way to resolve these difficulties for finite
systems (i. e., ferromagnetic particles which are finite by themselves or
systems where only a finite region creates the stray field-like certain kinds of
domain walls) is well-known. Namely, for such systems one can apply the
zero padding technique (often used in the signal processing) to avoid aliasing
due to the non-periodicity of the signal. For the evaluation of the dipolar field
the application of this technique means that we should: first, double the size of
the initial lattice padding the magnetization configuration with zeros in all
directions, and second, cut off the dipolar interaction matrices in Eq. (10. 18)
at the distance equal to the size of the initial lattice.
Then we can treat our system as having a periodic magnetization
configuration consisting of replicas of the initial system separated by areas of
the same size with zero magnetization. There is no interaction between
different replicas due to the cut-off of the dipolar matrices introduced exactly at
the initial system size. Hence, all physics of this artificial periodic
configuration are determined entirely by the interactions within one repl ica, as
it was for the initial system. On the other hand, both conditions of the
convolution theorem-the periodicity of the magnetization configuration and the
finite range of the interaction-are fulfilled and we can safely apply this
Numerical Simulation of Ouasistatic and Dynamic Remagnetization. . . 437

theorem to evaluate the dipolar field.


10.3.3.2 Periodic Boundary Conditions: FFT Combined with the Ewald
Method

As important as they are, finite systems with open boundary conditions do not
cover the whole range of physically interesting problems. Simulations of
magnetic systems which are so large that they should be treated as infinite to
avoid the influence of finite-size effects are necessary, e. g., when
investigating extended thin films, bulk magnetic materials, magnetic
nanocomposites, etc. The usage of periodic boundary conditions (PSC) is in
this case unavoidable and the question arises how to calculate the
demagnetizing field in such a situation.
First of all, it is evident that the direct summation Eq. (10. 17) cannot
beapplied at all, because we have to take into account the interaction between
all the replicas of the simulated area appearing due to the periodic boundary
conditions. Thus the number of terms in Eq. (10. 17) would be infinite.
The second idea is to introduce some cut-off for the magnetodipolar
interaction, thus taking into account the finite number of cells only, arguing
that this interaction decays quite rapidly (as (-3). However, this decay is not
fast enough to allow such a cut-off in 3D problems. It can easily be verified by
analyzing the behavior of the integrals from the dipolar interaction over, e. g.,
the spherical volume as the function of sphere radius R s: the integrals diverge
logarithmically with R s demonstrating that contributions from all distances are
equally important (which, by the way, enables the introduction of the
demagnetization factors for uniformly magnetized bodies of some simple
forms). The cut-off trick works, in principle, in 2D situations (simulations of
thin films) but even in this case one is forced to adjust the cut-off radius for
each new system to be studied to ensure that the errors introduced by the cut-
off do not affect the result.
The third candidate-the convolution theorem which saved the day for
open systems (see previous section)-cannot be applied for the dipolar field
calculations in systems with PSC. Although the magnetization configuration is
now periodic by the very definition of PSC, the interaction between various
repl icas of the simulated system is qual itatively important (in fact, this
interaction is the main reason to use such boundary conditions at all) and hence
the dipolar interaction cannot be truncated at any distance. This means that
the second condition necessary for the application of the convolution
theorem-the finite support of the convolution kernel-is violated.
The last hope in this situation is the Poisson equation
(10.20)
where p (r) is the density of artificial "magnetic charges" defined via the
magnetization M( r) as p (r) = - div M (r). This equation enables us (see
below) to establish the connection between the Fourier components of the
438 D. V. Berkov, N. L. Gom

magnetization configuration {M Cr)} and the magnetic potential . This


connection can then be applied to calculate or the stray field H diP = - grad
itself. However, the straightforward application of this method may lead to
disastrous results, as we shall demonstrate below.
To explain the principal difficulty arising in the Poisson equation method,
we shall apply this method directly to the lattice of point magnetic dipoles Jli
which charge density Pi can be expressed as Pi =Jli gradCoCrJ) CLandau
and Lifshitz, 1975). For a 2D N x x Ny lattice of magnetic dipoles the total
charge density is
Nx,N y

pCr) = pCrl1 ,z) =- oCz) ~ Jlij \J oCrll- rij) Cl0.21)


i.j

where the lattice plane is the Oxy-plane; the vector r II and vectors rij which
define the lattice node locations lie in this plane. For simplicity we also
assume that magnetic dipoles also lie in the lattice plane so that Jlij = /-I'ij ex +
/-It e y' For such a lattice with PSC the charge density Eq. C10. 21) is also
periodic in the Oxy-plane. Hence, using the expansion of Jlij and oCrij) into
Fourier series and oCz)-into the Fourier integral, we obtain the Fourier
expansion of the charge density as
+~

pCrl1 ,z) =- 2rr~S


-~
f dqzeiQzZ ~[/-IxCq//)qx + /-IyCq//)qy]eiQllrll.
{Qll}
Cl0.22)

Here I:lS = I:lx I:ly is the single cell area, /-I x(y) Cq II) are the Fourier transforms
of the dipole component arrays and the sum is taken over the wave vectors q II
corresponding to the simulated lattice.
Performing the same Fourier transformation for the magnetic potential ,
applying the Laplace operator to this Fourier transform, substituting the result
and the expansion Eq. C10.22) into Eq. C10.20) and using the orthogonality
properties of the Fourier harmonics, we obtain the Fourier components of the
magnetic potential as

C )- 4rri/-l x Cqll)qx+/-IyCqll)qy
cp q // ' q z - - I:l S q2 Cl0.23)

where q2 = q~1 + q;. For further calculations the Fourier components of this
potential in the 0xy-plane Cq II ' Z = 0) are required. They can be evaluated
by integrating Cq II ' q z) over q z leading to

-0)- 2rri/-l x CV/)qx+/-IyCqll)qy


Cl0.24)
C
cpqll'z- ---
I:lS qll

The Fourier components of the dipolar field H diP Cr) can be found using the
standard relation H diP Cr) = - gradC cp Cr. For the Fourier components it reads
Hx(y)Cqll)=-i' qx(y)' cpCq//) so that for HdipCq) one finally obtains
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 439

W(Q;I ,Z = 0) =- 2lT JJx(qll)q~ + JJy(qll)qxqy (10.25a)


b..S q II
W(Q;I ,Z = 0) =- 2lT JJx(qll)qxqy + JJy(qll)qt (10.25b)
b..S qll
According to Eq. ( 10. 25), the real-space stray field components at the lattice
nodes H xCy) (rij) (where magnetic moments are located) can apparently be
obtained in the following way: CD calculating the Fourier components
JJxCy) (qll) of the magnetization arrays JJij.xCy) using FFT ('" N 10gN
operations), CZ) multiplying them by the corresponding components of the
wave vector q II to obtain the Fourier transform Eq. ( 10. 25) of the stray field
H xC y) (q II) ('" N operations) and @ performing the inverse FFT to obtain the
real space components of H diP (again, '" N 10gN operations).
Equation (10. 25) looks exactly as the corresponding stray field
computation recipe for the open boundary conditions (Eq. (3.19)) : the Fourier
transform of the stray field is the product of Fourier transforms of the
magnetization configuration and the interaction kernel. However, for the PBC-
case this result was obtained in a completely different way (using the Poisson
equation). This derivation does not take into account some peculiarities of the
finite element method in contrast to the result (Eq. (3. 19)) for the open boundary
conditions and we will demonstrate now that the straightforward application of
Eq. (10. 25) does by no means guarantee correct results for systems with
periodic boundary conditions discretized into finite elements.
To illustrate this circumstance, we consider the simplest example of a
single dipole placed in the middle of the lattice (i = N x /2, j = Ny /2), with all
other sites empty. The answer for this case is well known: the scalar potential
of the stray field should be equal to the potential of a single point dipole
(shown in Fig. 10.6 as hollow circles), at least if we neglect the contribution
of the "replicated" dipoles appearing due to PBC; for a sufficiently large 2D
lattice corresponding corrections should be small everywhere except the
simulation area borders. The numerical answer obtained using the algorithm
for solving the Poisson equation outlined above (direct application of Eq. (10.25))
is shown in the same Fig. 10.6 as the solid line. This wildly oscillating function
has obviously very little in common with the correct behavior of the analytical
answer (open circles) .
This disaster is due to the following two circumstances. First, for a finite
periodic lattice we have at our disposal only a finite number of the Fourier
components. Second, the Fourier components of the scalar potential (Eq. (10.24))
and the corresponding field (Eq. (10. 25)) do not tend to zero for the large
values of the wave vector q (for the stray field they even diverge as '" q when
q-+-ClO). The first circumstance forces us to cut-off the Fourier spectrum of the
dipolar f.ield at the maximal available wave vector components q x'" 1/ b..X and
q y ' " 1/ b..y, where b..X and b..y are the single cell sizes. Due to the second fact this
leads to the sharp cut-off of the stray field Fourier-spectrum. Such a sharp cut-off
results, as it is well known, in large artificial oscillations of the field itself.
440 D. V. Berkov, N. L. Gorn

3
dip
Potential found from
2 the Poisson equation

Correct analytical result

-15 -10 10 15
o -I r (I.u.)

-2

-3
Figure 10.6 Potential of a single dipole placed in the middle of a 32 x 32 lattice: correct
analytical result (circles) and numerical result obtained using the Fourier components
(Eq. ( 10.25 obtained from the solution of the Poisson equation. See text for further
explanations.

In some cases this problem does not manifest itself because the variation
of the magnetization components along the lattice is slow (e. g., due to the
exchange interaction between the neighbouring cells). This means that both
J.Jij and J.Jt are smooth functions of the node position (i ,j). Hence their Fourier
components J.J(q//) for large wave vectors qx ....... 1/11x and qy ....... l1y are small
resulting in small Fourier components of the stray field (Eq. ( 10. 25. In this
case the cut-off of the stray field Fourier spectrum does not introduce any
substantial oscillations, because the spectrum components for large q' s are
already small by themselves. This is the reason why the direct evaluation of
the dipolar field using the relations like Eq. (10.25) can provide satisfactory
results (see, e. g., Refs. (Mansuripur and Giles, 1990; Giles et al., 1991.
However, in many systems such a nice behavior of the magnetization
components is not the case. For example, in polycrystalline extended films
rapid changes of the magnetization between the neighboring cells are possible
if the exchange across the grain boundaries is weakened. The second case
where such rapid changes are unavoidable is the simulation of the patterned
structures (arrays of nanowires or nanodots) where the magnetization
disappears by crossing the border between magnetic and non-magnetic
(empty) lattice cells. The problem is always present also in the simulations of
the thermodynamical properties of lattice models (Hucht et aI., 1995; Chui,
1996) where differences between adjacent magnetic moments are large at
least above the ordering temperature. In all these examples, the error
resulting from the sharp cut-off of the stray field Fourier-spectrum is
uncontrollable, and hence a method to resolve this difficulty is clearly needed.
The difficulty described above has been well-known for several decades:
Numerical Simulation of Ouasistatic and Dynamic Remagnetization. . . 441

it was noticed first when calculation of the Coulomb sums in ionic crystals was
performed (see, e. g., the textbook (Slater, 1967)). In such calculations the
divergence of the Fourier spectrum for the Coulomb kernel is even more
pronounced than for the dipolar case due to the slower decay of the Coulomb
interaction when compared with the dipolar one. The solution of this problem
for ionic crystals is known as the Ewald method, which can be adopted for the
dipolar interaction also (see Ref. (DeBell et ai., 2000) and Ref. therein). In
this section we briefly describe one of the simplest realizations of the method
(Berkov and Gorn, 1998); the version outlined below also allows the
straightforward generalization for systems of cells having finite and different
sizes in all spatial directions.
Following the basic idea of the Ewald method we add and subtract at each
lattice node (i ,j) (carrying the point dipole lLij) an artificial Gaussian dipole
with the charge density

_ 1 [ (r-rij)2]
Pij(r) - - (2lT)3/2
a
5 [(x - Xij)J.Jij.x + (y - Yij)J.Jij,r]exp - 2a 2 .

(10.26)

The choice of the "dipole width" a will be discussed later (here and below the
e5(z)-dependencies of the charge density are omitted for simplicity). One can
easily verify that the total dipole moment of the distribution (Eq. (10.26 is
- J.J ij' After this operation the charge density of the system can be written as
the sum of two parts,
(10.27)

where the first part

(10.28)

contains per) of the initial point dipoles system given by Eq. (10.21) and the
contribution of the "negative" Gaussian dipoles Eq. (10.26). The second part
y
1 Nx,N [ (r-rij)2]
Peer) =- (2lT)3/2
a
5~ [(x - Xij)J.Jij.x + (y - Yij)J.Jij.r]exp - 2a 2
/,}

(10.29)

represents the density of the" positive" Gaussian dipoles and is added to


cancel the second sum in Eq. (10. 28), thus restoring the initial charge
density.
The purpose of this addition-subtraction of the Gaussian dipoles is the
same as in the standard Ewald method (K ittel, 1953; Slater, 1967). First we
note that the total dipole moment of the first part PA (r) connected with each
lattice site is zero (at each lattice site the total moment of the negative
442 D. V. Berkov, N. L. Gorn

Gaussian dipole exactly compensates the point dipole moment of the initial
system). For this reason the field created by each lattice site due to PA (r)
rapidly tends to zero and may be treated as a short-range interaction. On the
other hand, the second part PB (r) of the charge density is smooth (it consists
of the smooth Gaussian distributions), so that the field created by this part can
be safely calculated using the Poisson equation technique (see the discussion
above) .
More precisely, the field created by the first part of the magnetic charge
density PA (r) associated with the lattice site (i, j) is

J..l ij ] f (Ll)
H;.ij (r) = [ 3 (x - X ij ) ( Jl ij Llr ij) -
( Ll r ij ) 5 ( Ll r ij ) 3 G r I}

~
_ _ (x - x)("Llr)
'I r"1 I} exp [(Llr.)Z
_ II ]
(10.30)
IT (Llr ij)5 20 2

(substitute y for x to obtain H y-components) where Llrij = r - rij and the


function f G (r) is defined via the standard error function erf( x) as

f G ( r) = 1 - erf (0 ~ )+ ~ ~ exp ( - 2~ 2 ) . (10.31>

Field Eq. (10. 30) decays for ro as-exp(- r 2 /20) and, hence by the
evaluation of the field H~P due to the part A of the charge density only the
contributions from several nearest neighbors of each cell (for which Llr ij - 0)
should be taken into account. This means that the H~iP -evaluation takes - N
operations for the whole lattice.
The field HgiP due to the second part of the charge density Eq. ( 10. 29)
can be calculated exactly as it was done for the .lattice of point dipoles
(Eqs.(10.20)-(10.25. The result is that the Fourier components of HgiP
can be obtained from the Fourier transform Eq. ( 10.25) of the field created by
the point dipoles by multiplying Eq. (10.25) by the factors exp( - q 1/02/2) .
These factors assure that HgiP (q II) tend to zero for large wave vectors so that
the spectrum cut-off due to the finite number of the Fourier components used
does not lead to any artificial field oscillations.
The last methodical problem is the choice of the Gaussian dipole width o.
From Eqs. ( 10.30) - (10.31) it can be seen that 0 should be chosen as small
as possible to ensure rapid decay of the short-range field H~iP, so that by its
calculation only the contributions from the few nearest neighbors should be
taken into account. On the other hand, 0 should not be too small, because
otherwise the factors exp( - q~1 0 2 /2) responsible for the decay of the large-q
Fourier components of HgiP will be not sufficiently small even for the largest
available wave vector components q~ax - 1/ Llx and q~ax - 1/ Lly. We found
that the choice 0 = max (Llx, Lly) for which the contributions from the three
nearest neighbor shells have to be taken into account by the H~iP-evaluation
provides the optimal compromise between the calculation speed and accuracy:
the relative error in Hdip due to the cut-off of the short-range part Eq. (l O. 30)
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 443

by the stray field evaluation was always less than 10-3.

10.4 Energy Minimization Methods

The choice of the best methods for the magnetic energy minimization is
crucially important for solving micromagnetic problems, because in order to
handle magnetic structures of practically relevant physical sizes one usually
has to discretize the simulation region in '" 104 - 10 6 cells. In addition, already
the nature of the static problems to be solved in micromagnetics is very
different for different systems (not to mention dynamic and thermodynamic
tasks, see Section 10. 6): there are problems including soft and hard magnetic
materials, localized and extended magnetization structures, bulk and thin film
geometries, etc. For this reason it is highly unlikely that the universal
minimization method to handle all micromagnetic problems will be ever found,
so we are forced to search for the best method for each specific problem. This
requires the knowledge of advantages and disadvantages of at least all
standard minimization techniques. In this section we give a brief review of
most minimization algorithms commonly used in micromagnetics and discuss
some new ideas concerning this topic.
As we shall see, the task of choosing the best minimization method is
very far from being accomplished. The only thing about which a general
agreement seems to exist is the termination criterion: in almost all modern
simulation packages the criterion based on the reduced torque is used.
Namely, the minimization is stopped when a torque acting on each magnetic
moment is smaller than some threshold: I (m i x h~ff) I < E. The values E =
10-3 - 10- 4 are usually small enough in the sense that no further changes in the
equilibrium magnetization state are observed if smaller E are used. This
criterion provides information about the state of each cell (particle) moment
and hence is more sensitive than the alternative criterion when the
minimization process is terminated if the total energy change during a single
iteration is sufficiently small.
Starting the discussion about the choice of the minimization method, we
recall first that, due to a complicated interplay of various physical
interactions, almost any magnetic system has many configurations (metastable
states) delivering energy minima to this particular system. Among them there
is obviously one global minimum, but it is by no means the only state being of
practical interest. For this reason we should first distinguish between the
algorithms searching for CD the nearest (to the starting magnetization
configurations) local energy minimum, CZ> the global energy minimum which is
unique for the system under study and should not depend on the state from
which we start the minimization.
444 D. V. Berkov, N. L. Gom

In general, the search for the global energy minimum should be performed
using the simulated annealing methods (for a simple introduction see (Press
et al., 1992. These methods are by their very definition quite slow so that at
present there is no sufficient information on how to apply them to large-scale
micromagnetic problems. For this reason we do not intend to discuss the
search of the global energy minimum of magnetic systems and draw our
attention to the problem of finding the nearest local (quasi) equilibrium or
(meta) stable system configuration.
To find such a local minimum, we can adopt one of the following
strategies:
( 1) Apply some standard numerical method for the minimization of many-
variable functions.
(2) Solve the equations of motion for the system magnetization (i. e., in
the Landau-lifshitz-Gilbert form). The dissipative term present in such
equations drives the system towards the equilibrium so that after a sufficiently
long integration time an equilibrium~ system state can be achieved with any
desired accuracy.
(3) Organize an iteration procedure based on the fact that in the
equilibrium state the magnetization should be aligned parallel to the
corresponding effective field. A simple alignment of each cell moment along
this field in each iteration does not work, as we shall see below, but it is
always possible to construct a closely related algorithm which surely
converges.
Below we analyze all three groups listed above. Our analysis is restricted
to those algorithms where the stray field is evaluated directly (using FFT,
FFT-Ewald or a direct summation). The latest development of the minimization
methods in systems where the magnetodipolar interaction is treated using the
Poisson equation (see Section 10. 2. 2) can be found, e. g., in (Gibbons,
1998: Yang and Fredkin, 1998).

10.4. 1 Standard Minimization Technique: Conjugate Gradients

The three most commonly used minimization techniques for many variable
functions are the steepest descent method, the method of conjugate gradients
and various quasi-Newton (variable metric) methods (Gill et ai., 1981;
Acton, 1990: Press et al., 1992).
To choose among these methods we should take into account the following
features of a micromagnetic problem:
( 1) The number of independent variables N is extremely large (-- 104 - 10 6 ) .
(2) We can calculate the energy gradient-the effective field He" =
- 8E IBM in the continuous formulation, which reduces to hT" = - <J E l<Jm; after
the problem discretization.
(3) In principle it is possible to calculate the Hessian matrix (matrix of the
Numerical Simulation of auasistatic and Dynamic Remagnetization. . . 445

second derivatives) of the energy also. However, the size of this N x N


matrix would be so large (see (1 that even its storage would be a real
problem-not to mention any operations with this matrix like its multiplication
by a vector or even inversion.
The first standard method mentioned above, the steepest descent method,
is known to converge extremely slowly due to its feature to perform many small
iteration steps going down along the long and narrow "valley." This behavior
is observed even for functions with relatively simple contour surfaces, so that
the usage of this method does not make any sense in micromagnetics.
Feature (3) from the above list also makes the usage of various quasi-
Newton methods unfeasible (Press et al., 1992): these methods, trying to
build up an approximation to the inverse Hessian matrix of the energy, require
at least ....... N 2 operations (for the matrix update) at each iteration.
Hence conjugate gradient methods (CGM) are the only reasonable
candidates among standard numerical methods for the function minimization.
They are used indeed by several groups involved in classical micromagnetic
(Williams and Dunlop, 1990) and fine particle (Fredkin et al., 1992; Berkov,
1996a) simulations and in some commercial micromagnetic packages (ARM).
However, these methods, when applied to micromagnetic problems, still have
serious disadvantages.
First of all we note that to implement conjugate gradients we must transfer
from Cartesian to spherical coordinates of magnetic moments (and to rewrite
corresponding formulas for the effective field and energy evaluation). The
reason is that during the remagnetization process, the magnitude of the
magnetic moment should remain constant (I MI = M s ) and hence its Cartesian
components are subject to the restriction M; + M~ + M~ = Const. This means
that these components are dependent variables which does not allow us to use
them by the implementation of CGM, because the construction of a set of
conjugate directions obviously fails when dependent variables are used.
Transition to spherical coordinates (em' 4>m) leads, however, to another
well-known technical problem (Nakatani and Uesaka, 1989; Williams and
Dunlop, 1990): if a particle moment is sufficiently close to the polar axis
(em"""'O or em"""'TT) then its small movements can lead to arbitrary changes of
4>m. If not properly avoided, these instabilities result at least in a very slow
convergence of CGM. In the worst case they lead even to a failure to property
build up a corresponding set of conjugate directions, so that the method gets
stuck at some magnetization state whose only fault is the unfortunate
orientation of some moments with respect to the arbitrary spherical axes
chosen by the user.
If only a single equilibrium state must be found, starting from the
configuration sufficiently close to this state, then the solution of this problem is
known (Williams and Dunlop, 1990; Berkov, 1996a; Ramstock and Dunlop,
1997). Namely, we can choose for each moment its own separate spherical
coordinate system with polar axis sufficiently far from the initial moment
446 D. V. Berkov, N. L. Gorn

direction. It is usually enough to choose between two different systems with


the polar axis along, say, z-and x-axis of Cartesian coordinates. In this case
the instability may not arise because during the minimization all moments will
remain sufficiently far from the polar axis of their coordinate systems.
The situation becomes much worse if we are interested in a sequence of
equilibrium states achieved, e. g., during the remagnetization process
(simulations of the hysteresis loop). In this case, the magnetic moment of
each cell can change its orientation by an angle--1T whereby the trajectory of a
this moment movement on the unit sphere may be arbitrarily complicated. This
difficulty forces us to implement the general conjugate gradient algorithm for
micromagnetic problems in the following way (Berkov, 1996a; RamstOck,
1997> .
First, having at our disposal some starting state, we choose for each
moment a corresponding spherical coordinate system according to the rule
introduced above (the polar axis should be as far as possible from the moment
direction). Then, at every iteration of the energy minimization process we
check whether it has approached a polar axis of its spherical coordinates for
each moment. If for any moment its polar angle em is less than a certain critical
angle ecr ' then the spherical coordinate system for this moment is changed and
the conjugate gradient method is restarted from the current state. This restart
is necessary, because after changing the coordinate system all information
collected about the conjugate direction set is useless CD .
Obviously the performance of the method constructed this way depends on
the critical angle value e cr used to decide whether the CGM should be
restarted. This value should not be too small to avoid the instabilities
described above. On the other hand, too large ecr-values would cause the
algorithm to restart too often, which leads to the loss of efficiency. We
investigated this question "experimentally" for a system of fine single domain
magnetic particles with the dipolar interaction (Berkov, 1996a) and found that
in this case the optimal ecr - value is::=::::::: 5 , but this value can by no means be
considered as a universal constant.
The conjugate gradient method with the modifications described above
can, in principle, be used for solving any quasistatic micromagnetic problem.
However, we point out in advance that the performance of CGM is significantly
decreased due to the restarts of the method needed to keep the moments away
from their polar axes. It was shown, at least in several important cases
(Berkov, 1996a; Ramst6ck, 1997; Lopez Diaz et aI., 1999) that CGM is
much slower than the equation of motion techniques and advanced relaxation

CD In principle, it might be possible to develop an algorithm which would transfer the


information already collected into this new coordinate system. However, such an algorithm is
expected to be rather complex due to the non-linear relation between the spherical
coordinates of a given point in different spherical systems.
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 447

methods described below. The discussion concerning the comparison of


different methods is postponed until the end of this section.

10.4.2 Equation of Motion Techniques and Simple Relaxation


Methods
Equation of motion techniques (and closely related simple relaxation methods)
are very popular both by solving classical micromagnetic problems (Bertram
and Schabes, 1988; Nakatani and Uesaka, 1989; Berkov et al., 1993; Yuan
and Bertram, 1991, etc.) and by simulating fine particle systems (Zhao and
Bertram, 1992; Vos et ai., 1993; Berkov, 1996a) due to their physical
transparency, guaranteed convergence and a large experience accumulated in
numerical analysis for solving systems of ordinary differential equations
(equations of motion).
The basic idea of all these methods is to build up a relaxation procedure
using equations of motion for a magnetization in an effective field:

dM;
dt =_ y[M x M eff ]
I I
- L[M x [M. x H eff ] ]
AM I , I ,
(10.32)
s

where i is the cell (particle) number and HIff is the effective field acting on the
corresponding magnetization. The constant y ( > 0) is equal to the
gyromagnetic ratio Yo in the small damping limit when the reduced damping
constant A --- O. The second term on the right-hand side in Eq. ( 10.32) is
constructed to account for the energy dissipation and hence the solution of
these equations should converge to the equilibrium magnetization configuration
(energy minimum) in the limit t --- 00. This means that after solving Eq. (10. 32)
for a sufficiently long time we are able to obtain the magnetization state
arbitrarily close to the equil ibrium(2).
If we are not interested in simulations of the magnetization dynamics, the
first term responsible for the precession of M; around the effective field can be
omitted, reducing the problem to the solution of the system of ordinary
differential equations (ODE)

~~; =- A[M; x [M; x HTff{M}]]. (10.33)

Here all constants have been absorbed into the reduced time T = tyM s and the
effective field dependence for each moment on the entire magnetization
configuration {M} is explicitly indicated.
There exists a huge variety of methods for solving ODE-systems like Eqs.

(2) By the way, Eqs. ('0. 32) - ( '0. 33) explain why the torque convergence criterion
I (m, x hr") I <E may be used: only the component of the effective field Hr" perpendicular to
the direction of the moment causes its movement.
448 D. V. Berkov, N. L. Gorn

(10. 32) - (10. 33). The major goal of the most sophisticated of them, such as
the predictor, corrector or Richardson extrapolation (Gear, 1971; Acton,
1990; Press et aI., 1992) , is to allow for a maximal time step maintaining
sufficiently high accuracy. In our case the task is much simpler: we do not
need to control the accuracy of our solution but merely have to assure its
stability. This means that we must only check that the system energy
decreases after each integration step.
Unfortunately, for most micromagnetic problems the equations of motion
are extremely "stiff" due to the competition between the strong short-range
exchange stiffness interaction and relatively weak, but long-range dipolar
forces. This means that even merely to keep the solution stable we have to
perform very small integration steps (typically ~ T ,..., 10- 4 - 10- 3 ) using
standard explicit methods for the solution of ODEs. And all established implicit
methods developed especially for solving such stiff ODE-systems require the
inversion of the system matrix for each time step. This inversion, being an"'"
N 3 -process for the N x N matrix, is absolutely out of question in our case (N
may be ,..., 10 6 ) so that the appl ication of these methods seems to be
impossible.
For this reason we have suggested that a version of the simplest possible
explicit method~the Euler method~may provide satisfactory results when
combined with a rather simple adaptive step-size control (Berkov et al.,
1993): if the energy after the integration with the time step ~T increases, the
magnetization configuration before this step is restored and the integration is
tried again with the time step ~ T /2. On the other hand, to avoid unnecessarily
small time steps we try to double ~ T if several (,..., 5 - 10) steps with the same
~ T have been successful (the energy decreased steadily) .
From the mathematical point of view, this algorithm is related to the
steepest descent method, because the rotation of the moment towards the
effective field direction corresponds to the movement in the local gradient
direction (keeping in mind that the magnitude of a moment should be kept
constant). An important advantage of the algorithm described above is that
only one function evaluation per iteration is required because we do not search
an energy minimum in the gradient direction. Even better, because we do not
perform a minimum search for each iteration, we may move closer to the
energy valley floor, than the standard steepest descent. This, in turn, means
that the moment movements for two subsequent iterations are not necessarily
perpendicular to each other, which also may lead to a faster convergence
when moving along a curved valley floor. The simple relaxation algorithm
outlined above performs surprisingly well, being at least in some cases much
faster than conjugate gradients (Berkov, 1996a; Lopez Diaz et al., 1999).
Further discussion is again postponed until the end of this section.
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 449

10.4.3 Advanced Relaxation Methods Combined with the


Extrapolation Techniques
Being confronted with a relatively poor performance of the standard
minimization techniques (especially the conjugate gradients) and equations of
motion approaches based on the known methods for the ODE solution, we
have tried to design our own minimization method for handling micromagnetic
problems. The resulting Modified Steepest Descent and Relaxation (MSDR)
algorithm developed by A. Hubert, K. Ramstock and the author was briefly
described in Ref. (Berkov, 1996a) and in more detail in the PhD work of K.
Ramstock (1997). For the sake of completeness we present here its short
description.
The algorithm consists of three stages:
(1) At the first stage several simple relaxation steps are made. The
simple relaxation step means that the system moments are rotated in the
effective field direction according to Eq. (10. 33) (thus moving the point
representing the system state in the direction of the local antigradient of the
energy E)
(10.34)

The step length a rei is chosen so that the energy value decreases. If after such
a simple relaxation step the energy increases, we return to the previous state
{m(k)}, decrease the step arel-a re l/2 and try again. The halving of arel is
repeated until it is small enough to ensure the decrease of energy-exactly as in
the simple relaxation method described above (see Section 10.4.2).
(2) If during several (typically -- 10) subsequent simple relaxation steps
with a constant a rei the energy steadily decreases, we perform the full energy
minimization in the local antigradient direction, i. e., we perform the move
Eq. ( 10.34) with the step length a rei chosen to minimize the energy value
E (let us denote the corresponding length as aE)' Usually aE is several
hundred times larger than a rei' which means that after several simple steps
(Eq. (4.5 with relatively small step length we are able to make a huge step
in the desired direction.
(3) The third stage is the minimization of the gradient norm of the target
function. In our case this means minimization of the reduced torque norm

(10.35)

along the gradient direction starting from the point where we have arrived after
the previous stage. It means that we perform the step Eq. ( 10. 34) choosing
a rei to minimize the sum Eq. (10. 35). The purpose of this stage is to
determine the step length for the subsequent simple relaxation steps (stage
450 D. V. Berkov, N. L. Gorn

one): a rei for the next simple relaxation steps is adjusted starting from a rei
obtained during this gradient norm minimization.
These three stages are repeated in the order described above until the
convergence criterion (max( I em; x hJffJ I )<E, see above) is satisfied.
For many problems (simulations of the equilibrium domain wall structures
in thin films (Ramst6ck, 1997), magnetic nanodots for MRAM applications
(Berkov et al., 2000), thermodynamical action minimization for transitions
between metastable states in fine magnetic particle systems (Berkov, 1998))
this method has provided the acceleration up to several hundred times when
compared with the standard steepest descent algorithm and the" equation of
motion" methods. It was also shown to be at least several times faster even in
comparison with the preconditioned conjugate gradient method (NAG-library)
(Ramst6ck, 1997) when the latter could be appl ied.
Such an acceleration is obviously due to the large step length aE during
the second stage (2) (the full energy minimization). Obviously, the rigorous
proof why should aE always be so large is hardly possible, because we know
too little about the energy landscape of the minimized functions (and we do not
claim that our method shall be the best in a general case). However, we can
try to explain why aE can often be much larger than the corresponding step of
standard minimization algorithms like the steepest descent. To do this, we
remember first of all that the steepest descent algorithm always (at each
iteration) tries to minimize the function in the antigradient direction (so that
each step is a full minimization step in our notation). This leads to well-known
oscillations across the energy valley because, if the starting point is on one of
the valley slopes, then the local antigradient vector is directed towards the
valley floor (not to the function minimum and not along the valley). Thus
performing the function minimization along this direction, one lands on the
opposite valley slope (and not on its floor, see, e. g., Ref. (Acton, 1990;
Press et a!., 1992)). These jumps between the two slopes may continue for
quite a long time, creeping slowly towards the function minimum along the
valley.
In our method we do not minimize the function along the antigradient
direction during the relaxation steps of the first stage. We merely choose arel
so that the energy E decreases. This leads to the step length arel usually less
than it would be if the full minimization along the antigradient direction is
performed. Hence, if we start from the point on the valley slope, after our
simple relaxation step we can land closer to the valley floor than after the full
function minimization. After several such steps we can finish so close to the
valley floor that the energy antigradient at our location is directed almost along
the valley (i. e., towards the desired energy minimum), thus allowing us to do
a really large step when performing the energy minimization along the
antigradient direction (the second stage) .
The third stage is necessary to adjust the relaxation step arel because
after performing the function minimization we can jump so far that the curvature
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 451

and the depth of the valley may change substantially. We note that although
the minimization of the gradient norm does not necessarily lead to the decrease
of the energy itself, it is almost always the case.

10.4.4 Alignment Methods


The underlying idea for these methods (which, by the way, were the first ones
used in the micromagnetics (Brown and La Bontc, 1965 is due to the basic
assumption that the magnitude of magnetic moments cannot be changed (the
material is assumed to be well below its Curie point) so that during the
minimization process only the moment rotation may occur. This means that to
reach an equilibrium it is sufficient to achieve the state where each magnetic
moment is parallel to the effective field at the corresponding point. In such a
state no further rotation of magnetic moments may occur, which can be seen
also from Eqs. ( 10.32) - (10.33): in this case all vector products on the right-
hand-sides of the equations of motion are zero.
This insight leads to a following "algorithm":
Step 1: AI ign each moment along the corresponding effective field Hrff .
Step2: For the state obtained, reevaluate Hrff everywhere.
Step3: Check whether the convergence criterion is met.
If not, accept this new state and go to Step 1;
If yes, we are done.
It is quite evident that the scheme based on (Step 1) - (Step3) only is
unlikely to succeed in practice in a general case (although some successful
applications of this algorithm were reported). The simplest way to understand
the reason for this failure is to realize that the algorithm sketched above is
nothing else but an attempt to solve a system of non-linear equations
mIX.; = e~.; {m} (10.36)
h
(e = Heffl H is a unit vector along the effective field, ex = x, y, z, and
= =
i 1, ' .. , N), substituting iteratively m~~i I) e~.; {m(k) }, where {m(k)} is
the magnetization configuration for the k -th iteration. It is a textbook result
where already in a 10 case the corresponding method for solving the equation
x = f(x) by simply putting x(k+1) = f(X(k) converges if and only if Idfldx I <
1 for the whole interval where we try to find a solution. There is absolutely no
reason why inequalities analogous to this one should hold in a general case for
Eq. ( 10. 36). Although the iteration process outl ined above can not diverge for
a micromagnetic problem (both mIX.; and e~,j are restricted due to the
condition m j = eli = 1), the algorithm can enter a non-convergent cycle, which
is what it really does.
The simplest solution for this problem is to use the underrelaxation
procedure: instead of aligning each moment along the effective field direction,
we update the moments according to the rule
452 D. V. Berkov, N. L. Goro

(10.37)
where the 0: j are called the underrelaxation parameters. After each iteration
according to Eq. (10. 37) all vectors mj are normal ized to preserve the
moment magnitudes. Making 0:' s sufficiently small we can reduce the rotation
angle for each moment, thus avoiding the cycl ic behavior of the algorithm and
assuring its convergence.
In standard micromagnetic problems we do not expect any significant
advantages of the algorithm based on Eq. (10. 37) when compared to the
simple relaxation employing Eq. (10. 33). First of all, if FFT-methods for the
stray field evaluation are applied, we are practically forced to use identical
underrelaxation parameters 0: j for all moments, because we have to move all
of them simultaneously (otherwise the gain achieved due to the application of
FFT is completely lost). Hence we have no information to adjust 0:' S for each
moment separately (such information could be obtained only analyzing changes
occurring after a single moment move). Second, due to the strong exchange
interaction present in such systems, small 0: -values are usually necessary to
avoid cycling mentioned above. And when all 0:' S are identical and small,
then the alignment method practically reduces to the simple relaxation
algorithm Eq. (10. 33), because for small 0:' s the move Eq. (10. 37)
(together with the subsequent moment normalization) leads almost to the same
result as the rotation Eq. ( 10. 33) .
The situation is completely different in systems where the stray field
cannot be evaluated using the FFT (at least at the present state of the art);
typical examples are disordered systems of fine magnetic particles, magnetic
nanocomposites, systems discretized using the irregular mesh. In this case the
sequential movement of single moments is not only possible, but even highly
desirable: first, the effective field can be re-evaluated after the movement of
each moment ("in place"), substantially decreasing the total iteration count;
second, the analysis of the changes taking place after a single moment rotation
enables the adjustment of the underrelaxation parameters 0: j for each moment
separately.
Such an analysis, performed for a disordered system of fine magnetic
particles with the dipolar interaction (Berkov, 1996a) enabled us to develop a
highly efficient (at least for this case) minimization method with the local
adjustment of the relaxation step. The information required for such an
adjustment was gained by analyzing the increments of each particular moment
during the iteration. This way the simple moment oscillations (with the period
of two iteration steps, where m~k) ~ mjk-ZJ) could be detected and the
relaxation parameter 0: j was adjusted to avoid them.
Although conceptually very simple, for the system studied in Berkov
(1996a) the method was shown to be several times faster than the conjugate
gradient algorithm and equation-of-motion method. The largest gain was
observed especially for small single-particle anisotropy values, where due to
the interparticle interaction, the cooperative remagnetization processes take
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 453

place, and hence the computation time required for the energy minimization is
particularly high.
After the description of each group of the minimization methods presented
above, one would expect the detailed comparison of the performance of all
these methods for typical micromagnetic problems. Unfortunately, such a
comparison is still not available. We know only three papers (Berkov, 1996;
Ramstock, 1997; Lopez-Diaz et aI., 1999) containing a systematic
comparison of various minimization methods at least for some particular
systems, but all these studies are incomplete.
In Ret. (Berkov, 1996) the disordered system of fine magnetic particles
with the dipolar but without the exchange interaction was studied. It was
shown that for this system the alignment method with the local adjustment of
the relaxation parameter is better than both the equation of motion technique
and CGM. However, it is evident that these results cannot be transferred to
the classical systems treated in micromagnetics because CD there exists a
strong exchange coupling in these systems and CZl in most cases the stray field
is evaluated using the FFT technique, which does not allow the local
adjustment of the relaxation step (see above) .
The PhD thesis of Ramstock (1997) contains the detailed description of
the newly developed MSDR method and its comparison with the preconditioned
CGM for several domain wall configurations. It is clearly demonstrated that
MSDR is superior with respect to CGM for all configurations studied
(especially for lattices with a large number of finite elements). Another
advantage of the MSDR is its much smaller memory requirements (as
compared to the large amount of memory required by the preconditioner
routines used in connection with CGM). But the MSDR method, although being
highly promising, is too new to allow any final conclusions.
Another attempt to compare various micromagnetic solvers was done in
(Lopez-Diaz et aI., 1999), where two typical problems-CD the search for the
equilibrium magnetization state for a rectangular Permalloy plate and CZl the
study of the Bloch wall interacting with a non-magnetic inclusion-were used to
analyze the performance of various solvers. It was found that for the first
problem the simple relaxation methods perform better than CGM (not to
mention the steepest distance), but for the second problem CGM was found to
be significantly (several times) faster than the simple relaxation.
Unfortunately, the implementation of the simple relaxation method in (Lopez-
Diaz et aI., 1999) seems to be incorrect (no adjustment of the relaxation step
during the minimization procedure was done), so that the results obtained
there should be revised at least at this point.
It is clear from this short discussion and the brief comparison presented
above, that the systematic and detailed study of various minimization methods
is still necessary to provide recommendation that could be used for the
majority of micromagnetic problems. Further development of new methods and
refinement of the existing ones are also highly desirable.
454 D. V. Berkev, N. L. Gem

10.5 Equilibrium Magnetization Structures and


Quasistatic Remagnetization Processes

10.5.1 Nanosized Magnetic Elements


Recent advances in numerical micromagnetics reviewed above have provided
powerful and reliable methods for computer simulations of quasistatic and
dynamic remagnetization processes in small thin film elements (with lateral
sizes in a sub-\-Im region). Such simulations are necessary, in particular, for
the development a new generation of computer memory-the so called
magnetic random access memory (MRAM) (Wang et ai., 1997; Tehrani et
ai., 1999). As shown above, numerical tools enable us to take into account
quantitatively the most important magnetic, geometric and structural features
of sub-\-Im thin film elements: their polycrystalline structure, distribution of
magnetic anisotropy axes of crystal grains, exchange weakening on the
intergrain boundaries, ferro- or antiferromagnetic coupling between different
magnetic layers, various element shapes, structural defects on the element
border, etc. In this subsection we show a few of our most interesting results
that may play an important role in optimizing practical applications of magnetic
nanoelements.
10.5. 1. 1 Single-layer Elements: Remanent States and Hysteresis Loops
One of the most challenging problems by constructing an MRAM-cell is the
optimization of the soft magnetic layer, i. e., the layer that should switch in a
relatively small magnetic field generated by the current pulses in the word and
bit lines. The current strength in these wires is strictly limited due to the
heating effects so that the maximal available switching fields are of the order
Hmax -100 Oe. This restriction requires a careful optimization of the shape and
other parameters of the soft magnetic layer in order to combine a sufficiently
high stability of the remanent state of this layer with a relatively low switching
field value. Below we show how this problem can be handled, taking layers
with relatively simple shapes as an example.
The first step is to find out which types of the remanent states for such
elements prepared from the soft magnetic materials are possible (starting from
the saturation). Performing corresponding simulations, we have confirmed that
all four well-known types of the remanent states (C-, C-inverse, $ and $-
inverse magnetization configurations-see Fig. 10. 7) can be observed for all
element shapes studied-squares, various rectangles with rounded corners
and elliptical elements. Which type of the remanent state is obtained depends
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 455

mainly on the particular realization of the crystallite configuration in the given


element.

C-dir C-inv S-dir S-inv

Figure 10. 7 Various types of the remanent state (shown as grey-scale maps of the
horizontal- m x - magnetization projection) for the magnetically soft rectangular element
with the lateral sizes a x b = 150 nm x 300 nm and thickness h = 2 nm (obtained starting
from the saturated state) .

The study of such remanent states is really important, because the


remanent state we start from is crucial for the switching mode of such soft
magnetic nanoelement. To show this, we have simulated the remagnetization
of such an element in a magnetic field of crossed wires with the geometry close
to that expected in real MRAM-cells (two perpendicular wires with the cross-
sections 300 nm x 300 nm, magnetic nanoelement placed in-between of them) .
For the switching starting from the S-state, an almost homogeneous
remagnetization takes place (Fig. 10. 8), whereby the remagnetization
starting from the C-state leads to the formation of complicated magnetization
structures including 360 domain walls (Fig. 10. 9). These walls form due to
the opposite initial deviations of moments (from the vertical Oz-axis) in quasi-
domains near the short element sides (Fig. 10. 7). The last type of a
switching process should obviously be avoided in MRAM-cells, so that the S-
type of the remanent state must be enforced somehow.
The next problem is the optimization of the lateral shape of the soft

D 0.5

-\'00 -{).75 -{).50 -{).25 0 0.25 0.50 0.75 1.00


H/Hmex
-0.5

I
Figure 10.8 The nearly uniform remagnetization of the rectangular soft (Py) magnetic
element starting from the remanent state of S-type (m x gray-scale maps). The element
sizes are 300 nmx 150 nmx2 nm; H max "':::450 Oe.
456 D. V. Berkov, N. L. Gorn

0.50 1.00
HIHmox

Figure 10. 9 Switching of the same element as shown in Fig. 10. 8 starting from the
C-state. Formation of a 360' domain wall is clearly visible: Hmax"",gOO Oe.

magnetic layer. Studying the remagnetization processes in rectangles and


elliptical particles with different aspect ratios, we could show, that, e. g., for
a rectangular element there exists an optimal ratio of the rectangle sides in the
following senses:
( 1) For rectangles with approximately equal sides the remanent state was
quite unstable already in small negative fields due to the formation of the
central domain (Fig. 10. 10).

WD
, - - - - - - - - - - , 1.0 MJM,

---r--'[S]-
-1.00 -{).50 0.50 1.00
H/Hmax

-1.0

Figure 10. 10 Remagnetization of a square soft magnetic element with rounded corners
(from the C-state); Hmax ""'450 Oe.

(2) For the rectangle with the side ratio a : b =


1 : 2, even for the
uniform remagnetization mode, the switching field H sw was too high for the
MRAM-applications (H sw ~ 230 Oe) due to the relatively large form
anisotropy.
(3) Rectangles with the side ratio a : b ~ 2 : 3 seem to represent an
optimal compromise having both sufficiently high stability of the remanent state
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 457

and relatively low switching field due to the moderate form anisotropy (H sw ""'"
130 Oe, see Fig. 10.11).
M.IM s
1.0

/\ ~I
,,-
0.5

-1. 00
'\ 0 0.50 1.0o
1-050

yO
H.IHmax
-0.5
.......
-1.0

Figure 10.11 The same as in Fig. 10.10 for the element with the side ratio
a: b=2': 3: Hmax~220 Oe.

Further we have also studied the effects of the homogeneous uniaxial


anisotropy induced in the soft element, influence of the average crystall ite
size, the intergrain exchange weakening and the boundary roughness on its the
switching behavior, etc. Corresponding results will be reported in detail
elsewhere.
10.5.1.2 Double-Layer Elements: Remagnetization of Artificial
Antiferromagnets
Another important problem of the MRAM is that development is the
stabilization of the so-called reference layer-the layer in which the
magnetization direction should be constant during the operation of the MRAM-
cell. One of the possible concepts of such a reference layer is based on the so-
called artificial antiferromagnet (AAF)-a double-layer element with the
opposite orientations of the layer magnetizations due to a strong interlayer AF
exchange coupling. This coupling should provide magnetization stability of the
AAF layer placed next to the switchable soft layer. The AAF-Iayer should be
" hard" enough to assure that its magnetization configuration is
CD homogeneous and (2) does not change substantially in the fields used for the
switching of the soft layer. Then the state of this AAF-Iayer can serve as a
reference magnetic state with respect to the soft layer by measuring the
corresponding changes of the resistance of this sandwich structure due to the
flip of the soft layer magnetization. It is clear that simulations of such an AAF
may provide useful insights into its expected remagnetization behavior.
Below we present one example of such simulations where typical
parameters expected for an AAF in MRAM-cells have been used: lateral size
a x b = 300 nm x 300 nm, equal layer thicknesses h = 1. 5 nm, spacer th ickness
d = 1.0 nm, magnetic parameters as for the cubic modification of Co
458 D. V. Berkev, N. L. Gern

(saturation magnetization M s = 1400 G, exchange stiffness constant A =


10- 6 erg/cm, cubic anisotropy constant K cub = - 1. 0 X 10 6 erg/cm 3 , 20
random distribution of the cubic anisotropy axes) , AF interlayer coupling J at =
0.4 erg/ cm 3 . A random crystallite structure inside each layer with the mean
crystallite size <0 > = 30 nm was generated using the method described in
Section 10.3.1.
The most striking result of our simulations is the existence of many
remagnetization modes for the same macroscopical initial conditions and
material parameters (Figs. 10. 12 and 10.13). For different realizations of the
random crystall ite structure of magnetic layers, with all other parameters
being completely identical, very different remagnetization processes were
observed. The variety of the switching modes found in our simulations covered
the whole spectrum from almost homogeneous (and hysteresis-free)
remagnetization (where magnetic moments of the upper and lower layers
simply rotate in the opposite directions due to the AF interlayer coupl ing, see
Fig. 10. 12) up to complicated remagnetization processes where even 360'
domain walls with different rotation senses inside a single wall were identified
(Fig. 10.13).

2 4
H./Ms

-1.0

Figure 10.12 Uniform switching mode of the double-layer AAF.

It is clear that only remanent states shown in Fig. 10. 12, namely those
with the almost homogeneous magnetization of both AAF layers, are allowed
for an MRAM-cell to be fully functional. Complicated domain structures like the
ones shown in Fig. 10. 13 would result in an unacceptable loss of the signal
amplitude (the latter is proportional to the difference of the CPP-
magnetoresistances between the states with two opposite orientations of the
magnetization of the soft layer placed above the AAF). Hence any domains in
the AAF-Iayers should be avoided. The simplest suggestion to enforce the
homogeneous magnetization rotation when preparing the remanent state of
AAF is to produce (in addition to the random crystallite anisotropy) uniform
uniaxial anisotropies in both AAF-Iayers. Corresponding anisotropy axes in the
upper and lower layers should be slightly tilted in opposite directions with
respect to the saturating external field. This would force the magnetization of
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 459

Figure 10.13 Complicate switching mode of the AAF with the formation of 360' domain
walls.

the upper and lower layers to rotate homogeneously inside each layer in the
opposite directions with respect to each other when the external field is
decreased, thus ensuring the uniform remagnetization mode.

10.5.2 Extended Thin Films and Patterned Structures


Simulations of extended thin magnetic films require the application of periodic
boundary conditions if the sample size in at least one of the lateral directions
exceeds several micrometers. In this case the major difficulty is the calculation
of the magnetodipolar interaction field, which should be found by combining
the FFT and Ewald techniques as explained Section 10.3.3.2. In all examples
listed below we have used this combination, refined to allow the handling of
finite size rectangular cells (in contrast to the point dipoles used in Section
10.3.3.2).

10.5.2.1 Hysteresis Loops and Ripple Structures in Extended Films


The conceptually simplest problem dealing with the remagnetization of thin
magnetic film is the magnetization reversal in an extended film occurring via
the formation of famous ripple structures (see (Hubert and Schafer, 1998) for
a recent review). The simulation procedure in this case is straightforward: we
start with the magnetization saturated in the initial direction of the external
field, gradually decrease, and then reverse this field.
A corresponding example is given in Fig. 10. 14. Here simulations of a
2.5 IJm x 5 IJm large area of a Permalloy thin film (using periodic boundary
conditions) were performed. The film thickness was d = 5 nm and an
additional homogeneous anisotropy in the Ox-(horizontal) direction was added
to the random crystallographic anisotropy to assist the ripple formation. It can
be seen that the ripple structure is already quite visible in the remanent state.
The coercive force (He::::';:: 10 Oe) is higher than normally observed for Py
460 D. V. Berkov, N. L. Gom

films, which may be due to the finite size effects and large average crystallite
size (0>""'='40 nm) chosen for the demonstration purposes.

-1.0

Figure 10. 14 Simulation example of thin film remagnetization via the formation of the
ripple structure (shown as m x -grey-scale maps) .

Using such simulations, the dependence of the microscopic ripple


structure parameters (like the correlation length in the direction perpendicular
to the external field and the characteristic wavelength in the direction parallel
to it) and the macroscopic parameters of the corresponding hysteresis loop
can be studied. A detailed example of such a study performed using the point-
dipole approximation for the stray field evaluation can be found in Berkov and
Gorn (1998).
10. 5. 2. 2 Critical Fields for the Domain Wall Motion
Another important problem which can be studied using the methods described
above is the determination of the critical field for the domain wall motion in thin
films. The examples of corresponding simulations are presented in
Fig. 10. 15a, b.
To determine the critical field we have first to create the domain wall
inside the simulation area. We do this by starting the simulations from the
magnetization configuration where magnetic moments in the right half of the
simulated region are aligned along the positive direction of the Oz-axis and the
moments on the left side have the opposite direction (see the rightmost image
in Fig. 10. 15a). This way we enforce the formation of the Neel domain wall in
the middle of the simulation rectangle. To prevent the formation of the second
wall with the opposite rotation direction on the vertical area borders, we set
the exchange coupling across these borders to zero.
After reaching the equilibrium in the zero external field starting from the
initial state described above we begin to increase the field in the direction
opposite to the Oz-axis, thus trying to move the domain wall to the left (the
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 461

Full intergrain_..,.-J-:-:- ---'-__H.+cr_....,.;::~rO-Q..,.......-.:.:=


exchange -<l.008

2.5 J.Ullx 1.25 ~m ........... O


o
""~---<::r (a)

Intergrain
exchange: IC'=O.I
--{).IO -<l.08 0.02

2.5 ~mx 1.25 ~m-----. _;;,..o.;:~

I---/.,...-------il II-I-...---i\1

Finite size effects Wall Wall pinned on the


jump intergrain boundaries
(b)

Figure 10.15 (a) Determination of the critical field for the domain wall motion in an
extended thin film with the full intergrain exchange; (b) The same as in (a) for the
intergrain exchange weakening K = O. 1.

right domain growths; the left domain vanishes). As long as the field is less
than the (quasistatic!) critical field of the domain wall motion, the
configuration remains stable (first three points in Fig. 10. 15a). But as soon as
the external field is able to overcome the wall pinning caused in this case by
the crystallite borders (due to the different directions'of the anisotropy axes in
different crystallites), the wall starts to move what can be seen as well on the
magnetization structure maps as on the remagnetization curve. \ The
corresponding critical field is marked in Fig. 10. 15a as Her and is in this case
Her~ 30e.
Imperfections on the intergrain boundaries are known to serve as strong
pinning centers. Indeed, assuming that the exchange across these boundaries
is 10 times weaker than the normal exchange interaction inside the grain, we
have observed a strong increase of the corresponding critical field
(Fig. 10. 15b) up to Her~ 25 Oe. The method also allows the study of the
Her-dependence on the grain size, anisotropy value and other magnetic and
462 D. V. Berkov, N. L. Gorn

geometric parameters of the film.

10.5.2.3 Remagnetization of Patterned Magnetic Structures


Simulations of patterned magnetic structures also require the application of
periodic boundary conditions, especially if the interaction betWeen the
structural units are thought to be important. The simplest case of such a
structure is an array of single-layer nanodots, which may have different lateral
shapes. Below we present simulation results for an array of circular nanodots.
In both cases shown below in Fig. 10. 16a, b the simulation area
contained an array of 4 x 4 nanodots. The total size of the simulation area

-{).4 0.2 0.4


H,IMs

(b)

Figure 10.16 (a) Remagnetization process in an array of separated circular Permalloy


=
nanodots. Dot diameter is 0 500 nm, the distance between the dot centers A 640 nm,=
dot thickness h = 10 nm. The absence of any correlations between the magnetization
structures of different dots is clearly visible; (b) The same as in Fig. 10. 16a, but for the
closely placed nanodots (0 = 11 = 640 nm). The collective nature of the remagnetization
process is evident.
Numerical Simulation of Ouasistatic and Dynamic Remagnetization. . . 463

was=5 IJm x 5 IJm and this area was discretized into 256 x 256 cells. The dots
were assumed to consist of Permalloy with the dot thickness h = 10 nm and
other parameters as indicated in the figure captions.
The main difference between the cases shown in the two figures is due to
the interdot interactions. For the separated dots (Fig. 10. 16a) with the
smallest distance between the dot borders being about= 140 nm the interaction
between different dots plays no role at all, which can be recognized by
considering the magnetization structures of various dots: they are totally
uncorrelated, because the interdot interaction is negligible (the interdot
spacing (140 nm) is much larger than the dot thickness (10 nm. Such a
remagnetization occurring via the individual dot switching leads to a relatively
broad hysteresis loop (Fig. 10. 16a) .
In contrast to the separated dots shown above, the system of closely
placed dots (Fig. 10. 16b) demonstrates a switching process dominated by the
cooperative behaviour of the array units: magnetization structures of all dots
are nearly the same and the resulting hysteresis loop is much narrower than for
the individual switching.

10.5.3 Quasistatic Remagnetization in Nanocomposites:


Individual Particle Switching and Cooperative
Remagnetization Processes
Magnetic nanocomposites are a class of advanced magnetic materials
consisting of magnetic nanoparticles embedded into a matrix with magnetic
properties different from those of particles themselves (including a non-
magnetic matrix). Their well-controlled production has become available
during the last decade. Currently such materials are a subject of growing
interest due to their great technical potential: they are believed to be the most
promising candidates for designing permanent magnets (Hadjipanayis, 1999),
interesting candidates for the development of sensors using the GMR-and CMR-
effects (see, e.g., Ref. (Altbir and Vargas, 1998: Kuzminski et aI., 1999
and the first choice for high density particulate recording media (0' Grady,
1999) .
Numerical simulations of magnetic nanocomposites lie somewhat aside
from the main road of numerical micromagnetics. On one hand, simulations of
nanomagnets consisting of two magnetic phases are more difficult than
standard micromagnetic simulations of homogeneous materials (Schrefl et aI.,
1994a, 1994b). On the other hand, in nanocomposites, where magnetic
particles are embedded into a non-magnetic matrix, the exchange interactions
between these particles are absent. This leads to a completely different
behavior of such systems making them closer to the spin glasses than to
"normal" ferromagnets.
However, as it should be clear from the impressive list of the potential
464 D. V. Berkov, N. L. Gom

applications of nanocomposites given above, such simulations are very


important from the point of view of applied physics. Due to the similarity
between certain kinds of such composites and the spin/dipolar glasses,
corresponding research is also very interesting from the fundamental point of
view.
In this subsection we will briefly discuss numerical simulations of
quasistatic remagnetization processes in the interacting Stoner-Wohlfarth
model-a disordered system of single-domain magnetic particles with the
uniaxial anisotropy and a magnetodipolar interaction between them. Results
for the remagnetization dynamics are presented in Section 10.6.
Simulations of the quasistatic remagnetization behavior of such systems
can be performed using all the methods discussed in Section 10. 4, whereby
due to the spatial-the FFT methods for the calculation of the dipolar field can
not be applied (at least not directly, see Ref. (Deserno and Holm, 1998) for
a detailed discussion). For this reason the alignment methods with the local
adjustment of the relaxation parameters (see Section 10.4.4) seem to be the
best ones, having decisive advantages over both the conjugate gradients and
the equation-of-motion algorithms. In Berkov (1996a) we have reported
simulations on an antique HP-712/60 MHz workstation on systems having up to
4 x 10 3 particles, which was more than enough to keep statistical errors for all
physical quantities of interest in frames of a few percent (in the worst case) .
For a disordered fine particle system there also exists an alternative (with
respect to the "true" numerical simulations) possibility to obtain
remagnetization curves- the so-called random field approximation (RFA). In
this approximation any physical quantity for a system of interacting particles is
evaluated as a convolution of the corresponding quantity for an ideal (non-
interacting) system and the distribution density of the interaction field p (hint) .
For example, the magnetization as the function of the external field h o can be
calculated as

m(ho. z ) = fm CO
) (h) p(h-ho)dh (10.38)

where mCO) (h) denotes the magnetization of an ideal system.


Before pointing out the obvious advantages of this method, we emphasize
already here the major drawback of RFA: in this approximation all interparticle
correlations (in our case correlations between particle magnetic moments) are
neglected. Indeed, according to Eq. (10.38) the interaction field on each
particle hi;nt is considered as a random variable with some a priori known
distribution density. Hence h\nt on the given particle is not correlated with the
orientations of its moment or neighboring moments. For this reason we expect
RFA to fail for systems where collective remagnetization processes are
important, i. e., systems with low single-particle anisotropy values and
moderate and high concentrations of the magnetic phase.
The big advantage of the RFA is that the calculation of magnetization
Numerical Simulation of Quasi static and Dynamic Remagnetization. . . 465

curves in this approximation are very simple and fast: we need merely an
effective and reliable method for the evaluation of the integral Eq. (10.38) for
given system parameters. Obviously, this integral can be evaluated only
numerically, because both multipliers in the integrand - the ideal magnetization
dependencies m(O) (h) and the distribution density of the interaction field
p (hint) -can be obtained only numerically even for the Stoner-Wohlfarth
model. However, this is a minor problem: first, the ideal magnetization curve
can be calculated once and for all with any desired accuracy. Second, to
generate quickly the interaction field with the desired distribution p (hint) the
following straight-forward procedure may be applied (Berkov, 1996).
As for real simulations, a random but non-overlapping particle
configuration with randomly oriented anisotropy axes is generated. Using the
magnetization state obtained in the previous field value (saturated state for the
first field), we evaluate the interaction field on each particle using, e. g., the
improved Lorentz cavity method. Then these interaction fields are
"redistributed" randomly between particles: the dipole field acting on the i -th
particle was set to h~nt, where j '#- i is chosen randomly. This way the
interaction field distribution corresponding to the given orientation degree of
magnetic moments is generated and the field acting on each particle is not
correlated with its position and magnetic moment orientation, as required by
theRFAEq.(10.38).
After generating all interaction fields, an equilibrium moment orientation
for each particle can be found using the alignment method described above. It
works in this case extremely fast because we do not need to re-evaluate
interaction fields during the iteration process. In principle, corresponding
equilibrium orientation angles can be tabulated even in advance (for the given
anisotropy value) as a function of the anisotropy axis orientation and the total
field (which in this case would be a sum of the external field and random
interaction field). This table can then be used to evaluate the equilibrium
moment orientation for any anisotropy axis direction and interaction field value
encountered by simulations by a simple interpolation. This trick would result in
further acceleration of RFA calculations, but we did not use it because the
computational effort to obtain hysteresis loops with accuracy better than 1%
using RFA was negligible even when the iterative alignment procedure was
used.
The methodologically most relevant question concerning such simulations
is the validity region of the RFA, i. e., for which particle anisotropies f3 and
volume concentrations c does the RFA provide adequate results. To answer
this question, we have performed simulations for systems with single-particle

The behavior of the interaction field distribution with the growing particle
concentration is fairly complicated, changing from restricted Lorentzian for small particle
volume fraction to nearly Gaussian for large concentrations CBerkov, 1996b).
466 D. V. Berkov, N. L. Gorn

anisotropy constants O. 0~f3~ 10.0 and particle volume fractions 0.0 1~ c~O. 32
using RFA and compared them with the results of "true" simulations carried out
using the alignment method (i. e., where the actual interaction field evaluation
for each iteration was used). As a measure of the error introduced by the
RFA, we have chosen an integral difference between hysteresis loops
computed using RFA m~FA and "real" simulations m~m, normal ized on the
hysteresis loop area ShYS. :

8 = -sl
hyst
f I m~FA(hz) - m~m(hz) I dh z . (10.39)

For those (f3, c)-values where this quantity is small (8 1) RFA can be
considered as valid and used by the numerical treatment of quasistatic
remagnetization processes as a much faster alternative to "real" simulations.
Contour lines on the (f3, c)-plane corresponding to the difference values
8 = 0.02, O. 1, 0.2 are shown in Fig. 10. 17 (note the logarithmic scale of the
f3-axis). If we consider a relative difference of 10% (8 = O. 1) as acceptable
then the dashed area on the Fig. 10.17 represents the RFA validity region. It
can be seen that, as expected, for higher particle concentrations larger
anisotropy values are required to make the RFA valid.
10

---6- 0=0.02
- 0 - 0=0.1
-v- 0=0.2

0.1 '-- -'- -----'- . L -_ _

o 0.1 0.2 0.3


c
Figure 10. 17 Phase diagram in the anisotropy-concentration coordinates showing the
validity region of the random field approximation (dashed area). The diagram is based on
the integrated relative difference {j defined in Eq. (10. 39) between hysteresis loops
calculated using RFA and "real" simulations. Above the solid line corresponding to {j = 0.1
this difference is small enough to justify the RFA usage.

This diagram also has an important physical sense because it can be


viewed as a "phase diagram" showing the regions on the anisotropy-
concentration plane where either single-particle or collective remagnetization
processes dominate the quasistatic behavior of the system. Indeed, the main
RFA assumption that the interaction field is not correlated with the particle
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 467

position and its magnetic moment orientation means that all collective
remagnetization processes are neglected. Hence, the failure of the RFA
signals the establishing of the collective remagnetization behavior, so that the
solid line in Fig. 10.17 can be considered as a "phase boundary" on the (13,
c) -plane. Above this boundary the remagnetization processes are dominated
by the single-particle, and below it, by the collective behavior. Only below
this line the collective spin-glass-like behavior (like a phase transition to a
"frozen" state) can be expected.
From the point of view of potential applications hysteresis loops of such
systems are of a major interest. A typical set of hysteresis loops for various
anisotropy values (for a particle concentration c = O. 16 as an example) is
shown in Fig. 10.18.

- P=IO
-0- P=3.2
- - - - P=I.O
----<>- P=0

-2 -I 2
H,IMs

-1

Figure 10. 18 Hysteresis loops for the particle concentration c = O. 16 and various single-
particle anisotropy values f3 as shown in the figure. Note the hysteretic behavior for the
system of isotropic particles (f3 = O. 0) - see text for detai Is.

It can be seen, that for the high anisotropies 13 -- 5 - 10 the loops for an
interacting system are close to ideal hysteresis loops for a non-interacting
Stoner-Wohlfarth model (Wohlfarth and Stoner, 1948). For such high
anisotropy, practically the only effect of the interparticle interaction is to
remove the square-root singularity in the static magnetic susceptibility of the
corresponding non-interacting system (Berkov and Meshkov, 1988). For
intermediate anisotropy values, interaction effects gradually became more
important, and finally, for small anisotropies these effects completely
determine the hysteretic behavior of the system. In particular, for 13<0.3
hysteresis loops do not depend on the anisotropy constant value at all and
practically coincide with the loop for isotropic particles (13 = 0, see
Fig. 10. 18). From the fundamental point of view this feature-the existence of
the hysteresis in a collection of fully isotropic particles- is quite interesting,
arising solely due to the collective interparticle interaction effects. It was first
predicted in Berkov and Meshkov (1990) and its existence was later confirmed
468 D. V. Berkov, N. L. Gorn

in Fredkin et al. (1992) and Zhao and Bertram (1992).


The two most important parameters of the hysteresis loop-reduced
remanence j R = m z ( h z = 0) and coercivity he are shown as functions of the
particle anisotropy {3 and particle volume concentration c in Fig. 10.19.
0.6 -+- (3=5.6 0.8 -0- (3=5.6
- 0 - (3=1.8 --4- (3=1.8
0.5 t1P'1il=_- ---4_ --0- (3 =0.56 --0- (3 =0.56
-+- (3 =0.18 ----T- (3 =0.3 2
--0- (3 =0 --0- (3=0
.-2; 0.3

0.2
0.1

o 0.1 0.2 0.3 o 0.1 0.2 0.3


c c
(a) (b)

Figure 10.19 Concentration dependencies of the remanence jA (a) and coercivity H e / Ms


(b) for various anisotropy values indicated on the figures.

It can be seen that for large anisotropies the remanence jA depends on c


only slightly (Fig. 10. 19a) and its value is very close to the remanence
j~O) = 0.5 of a non-interacting system of particles with a uniaxial anisotropy and
a chaotic distribution of easy anisotropy axes. For very low anisotropy values,
the jR ( c) dependence approaches the limiting curve for {3 = 0 when the
concentration increases (Fig. 10. 19a).
The coercivity for large anisotropies also depends on the particle
concentration only slightly and is again very close to the ideal Stoner-Wohlfarth
value h e( He/Ms)~0.48{3 (Wohlfarth and Stoner, 1948). The decrease of
he with growing particle concentration for large anisotropy values results from
the influence of the effective field fluctuations due to the interparticle
interactions. Such fluctuations always reduce the number of metastable states
in a system if these states result from single-particle effects (Ziman, 1979).
The most interesting feature of the coercivity concentration dependence is
its non-monotonic behavior for systems of particles with small anisotropies,
where he is determined entirely by the interaction effects (see the
corresponding curve for (3 = 0 in Fig. 10. 19b). The coercivity maximum near
c~O. 15 is quite flat but clearly beyond statistical errors. A possible
explanation of this behavior can be given as follows. When the particle
concentration increases, starting from very small values the interaction effects
responsible for the hysteretic behavior (by small single-particle anisotropy
values) become more pronounced leading to the initial increase of he (c)-
dependence. When the concentration increases further, a short-range spatial
order is formed (Ziman, 1979) so that a degree of disorder for larger
concentrations is reduced. This leads to the decrease of he: it is well known
Numerical Simulation of Ouasistatic and Dynamic Remagnetization. . . 469

that the hysteresis is absent for a fully ordered particle system, at least for
cubic and hexagonal lattices.

10.6 Equilibrium and Non-Equilibrium Thermodynamics:


Langevin Dynamics, Monte Carlo Method and
Path Integrals
We now turn our attention to simulations of the dynamic remagnetization
processes tak ing into account thermal fluctuations, i. e., to the study of the
equlibrium and non-equlibrium thermodynamics of magnetic systems.
In our contribution we are not going to discuss studies of the
remagnetization dynamics without temperature fluctuations. This problem,
although technically quite difficult, reduces formally to the integration of the
corresponding equations of motion (in most cases the Landau-Lifshitz Eq. (10.32) is
used) and the only task to be treated is the choice of the optimal method for
the solution of these equations. Numerical analysis has accumulated, as it was
already mentioned in Section 10. 4. 2, a nearly exhaustive experience in
solving systems of ordinary differential equations (see e. g., Refs. (Gear,
1971; Press et al., 1992), etc.). Of course, there exist several serious
problems specific for numerical micromagnetics; the interested reader may
consult the recent, very detailed and also pedagogically excellent introduction
to this area by Miltat et a!. (Miltat et a!., 2001).

10.6.1 Fast Remagnetization Processes: Langevin Dynamics vs.


Monte Carlo Method
10.6. 1. 1 Langevin Equations as the First Choice for Short Times and
Significant Precession Effects
Straightforward simulations of the remagnetization dynamics taking into
account thermal fluctuations amount, in contrast to the previously mentioned
topic, to the solution of the so-called stochastic differential equations, which is
by far a more difficult and a highly non-trivial task.
To be more specific, we restrict ourselves to the usage of the stochastic
Landau-Lifshitz-Gilbert equation (Brown, 1963b) for the magnetic moment
motion in its standard form

deft; =- Y[Jl; x (m + HD] -


ff
i\ : {Jl; x [Jl; x (m + HD]}.
ff

(10.40)

Here the precession constant Y is equal to the gyromagnetic ratio Yo in the


470 D. V. Berkev, N. L. Gem

limit of small damping A 1. The deterministic effective field H~ff acting on the
i-th magnetic moment includes all the contributions mentioned in Section 10.1
(external field. anisotropy field. exchange field and magnetodipolar
interaction field). The fluctuation field flI;' will be discussed below.
First. we would like to point out that we will not discuss here whether this
equation can be considered as a generally best choice. because this problem
is far beyond the scope of this chapter. For the same reason we cannot
address the question of how to chose the damping parameter A. We only point
out that this parameter accumulates both the information about various energy
dissipation mechanisms in magnetic systems (Suhl. 1998) and effects resulting
from the finite-element approximation of the continuous magnetization field
(Feng and Fischer. 2001). For these reasons A depends. generally speaking,
both on the characteristic time and length scales involved in each specific
problem. Hence, it is not allowed to use the damping value obtained from.
e. g . the FMR data for simulation of. e. g.. dynamics of the domain wall
motion. An interesting recent discussion concerning this question can be found
in Thiaville et al. (2001).
We now focus our attention on thermal fluctuations which are included into
the description of the magnetic moment motion via the so-called fluctuation
(Langevin) field HI! in Eq. (10.40). It is usually assumed that its Cartesian
components are o-correlated in space and time (Brown. 1963b)

(H~.;) =0
(HIL(O). Hf~.j(t) = 2Do(t)o;jOs'l' (10.41)

(here i. j are the moment indices and ~, \jJ = x, y. z). The noise power 0
can be evaluated using the fluctuation-dissipation theorem and is proportional
to the system temperature T and depends on y and the damping constant A :

o = _A_ kT
1 + A2 Y/J
Using the unit vector of the magnetic moment m=/J/(M s I:::.V) (where M s
denotes the material saturation magnetization and I:::. V is the volume of a
discretization cell or a fine particle if a disordered system of single-domain
particles is studied). the reduced magnetic field h = H/ Ms. and the reduced
time T= tyM s ' we can rewrite Eq. (10.40)as

~~; =- [m; X (h~ff + h'/)J - A{m; x [m;(h;ff + h'/)J}. (10.42)

The problem which arises when solving the Eq. (10.40) (or Eq. (10.42-
and many groups involved in numerical micromagnetics are still unaware of-is
that these equations can by no means be interpreted as usual differential
equations. This fact can be easily demonstrated using the simplest analogue to
Eq. ( 10.40) -the equation of motion for a single particle in a viscous medium
in the presence of thermal fluctuations
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 471

(10.43)

Here 1) denotes the friction coefficient dependent on the particle geometry, the
fluid viscosity, Fdet is the deterministic force acting on the particle, and ~L is
the reduced random (or Langevin) force. It is normally assumed to be a
random variable with the Gaussian distribution and correlation properties
analogous to those of the random field projections Eq. (10.41) :
<~(O> = 0, <~ (0) ~( 0 > = 205 Ct) (10.44)
with the fluctuation power 0"" T. The "good" function a (x, 0 is introduced to
show, e. g., that the noise characteristics may be coordinate- and time-
dependent.
Now, an attempt to integrate Eq. (10. 43) as a usual differential equation
unavoidably leads to the integral of the type
t

W(O = f~(ndt'.
o
This integral obviously represents some kind of a random process, because its
integrand is a random variable. From the correlation properties (Eq ( 10. 44
it can be easily deduced that W( 0 is the so-called standard Wiener process,
better known among the physicists as the Brownian motion (Gardiner, 1997).
The unpleasant feature of this process is that it is not differentiable - the ratio
[W ( t + 11 t) - W ( t) J/11 t diverges in the limit 11 t-O almost surely (the latter
does not mean, that there still exists a small hope that this limit is finite;
consult any handbook on stochastic processes for the corresponding exact
definition). Because the derivative dW/dt = ~(O does not exist, in the usual
sense Eq. (10. 44) including this derivative does not exist also and hence
cannot be interpreted as a "normal" differential equation.
There exists, fortunately, a possibility to assign a correct meaning to the
relations like Eqs. (10.43) and (10.40). The heuristic way to do this is to
rewrite the product ~ ( t) d t as the differential of the Wiener process ~ ( t) d t =
d Wand to try to define the integrals containing this differential

1= fa(X,OdWCt)

analogously to the standard Riemann-Stieltjes integrals as the limit of


corresponding partial sums
n
I = lim 2: a[xCT;), T;][I1W(I1t;)J
n- oo i= 1
n

= n_limL.:a[xCT;),T;][W(t;)
oo ;= 1
- WCt i - I )]. (10.45)

The points T i where the values of the integrand should be evaluated lie, as
472 D. V. Berkov, N. L. Gorn

usual, somewhere inside the interval [t ;-1' t; J.


The first problem is how to define this limit. In stochastics one has - in
contrast to the standard analysis-four possibilities (see, e. g., Ref.
(Gardiner, 1997), Chapter. 10. 2. 9); fortunately, one of them-the limit
defined in the mean square sense-is convenient enough to develop the
complete analysis of such stochastic integrals. The second problem is that this
limit itself-and not just the values of partial sums in Eq. (10. 45)-depends on
the choice of the intermediate points T; (see (Gardiner, 1997), Chap. 10. 3
for a simple but impressive example). This fact is in a heavy contrast to the
standard analysis where the independence of the limit of Darbu sums on the
choice of intermediate points is one of the cornerstones when building the
integral calculus.
The only way to cope with this second problem is to introduce some
standard choices of the intermediate points and to find the best choice from the
point of view of the statistical physics. Currently there exist two such standard
choices:
( 1) The intermediate points T; coincide with the leftmost points of the
interval: T; = t ;-1; in this case we speak about the Ito stochastic integral.
(2) The intermediate points T; are in the middle of the interval: T; =
( t ;-1 + t;) /2; in this case we obtain the Stratonovich stochastic integral.
It is well-known that the Ito and Stratonovich interpretations of a
stochastic equation lead to different solutions if the noise in this equation is
multiplicative-i. e., the random term is multiplied by some function of the
system variables. In this case usually the Stratonovich interpretation provides
physically correct results, recovering, e. g., some important properties of
physical random processes obtained using more general methods (Gardiner,
1997).
The noise in the Langevin Eq. (10. 42) (or Eq. (10. 40 is obviously
multipl icative, because due to the vector products the projections of the
random field H I' are multipl ied by the magnetic moment projections. This fact
was noticed already in the pioneering paper of Brown Brown (1963b) who
suggested that the Stratonovich interpretation of the this equation should be
used. For quite a long time afterwards the question was abandoned because
analytical solutions of Eq. (10.42) are available only in a few of the simplest
cases and computers were not powerful enough to enable numerical studies of
really interesting magnetic systems.
During the last decade, however, corresponding numerical simulations
became available and many research groups have performed the studies of
remagnetization processes in various systems using solving Eq. (10.42) -see
(Garcia-Paliacios and Lazaro, 1998; Zhang et ai., 2000; Berkov et ai., 2002;
Scholz et ai., 2001; Lyberatos and Chantrell, 1993; Nakatani et ai., 1997)
For such simulations the question concerning the choice of the stochastic
calculus is of primary importance, because different numerical methods
converge to different kinds of stochastic integrals. The Euler scheme and the
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 473

simple implicit methods obviously converge to the Ito solution, Heun and
Milstein schemes are known to converge to the Stratonovich limit (McShane,
1974) and Runge-Kutta schemes can converge to anything (including the in-
between cases) depending on the coefficients used there (Ruemelin, 1982).
Most authors (Garcia-Palacios and Lazaro, 1998; Berkov, 2002; Scholz et
aI., 2001) and commercial micromagnetic packages (ARM, LLG) use the
Heun and Runge-Kutta methods (simply because they are by far more stable
than the Euler method) converging to the Stratonovich solution, but several
groups employ the Ito-converging Euler (Zhang, 2000; Lyberatos and
Chantrell, 1993) and implicit schemes (Nakatani et aI., 1997). The last two
papers were seriously criticized in (Garcia-Palacios and Lazaro, 1998) where
it has been claimed once again that only the Stratonovich interpretation
ensures the physically correct solution of Eq. (10.42) so that results obtained
with methods converging to the Ito integrals should be discarded as incorrect.
In this subsection we shall prove analytically-and support our proof with
numerical experiments-that if the time evolution of the system is governed by
the stochastic Eq. (10. 42) then there is no difference between the system
behavior for the Ito and Stratonovich interpretations of this equation. The only
additional condition is that the magnitude of magnetic moments is assumed to
be constant@. This is the case in the overwhelming majority of models
developed for the description of magnetic systems, e. g., in the classical
Heisenberg model, in models describing RKKY spin glasses and fine magnetic
particle systems (Dotsenko, 1993; Hansen and morup, 1998) and in standard
micromagnetics (Brown, 1963a) (the last example represents probably the
most relevant research area from the practical point of view) .
First of all, we note that the fluctuation field in the dissipation term on the
right-hand side of Eq. (10. 42) can be omitted; although the particular
realizations of the system trajectories will be different then, the average
system properties (which are the only being of practical interest) remain the
same if the noise power 0 is rescaled correspondingly-see, e. g., (Garcia-
Palacios and Lazaro, 1998; Braun, 2000). Thus we can restrict ourselves to
the study of a simpler equation

~~; =- [m; X (hI" + hDJ - i\[m; x (m; X hI")]. (10.46)

The most straightforward way to show why the multipl icative noise in Eq. (10. 46)
does not lead to any difference between the Ito and Stratonovich
interpretations of this equation is to analyze the additional drift term appearing
by the transition between the Ito and Stratonovich forms. Namely, it is well-

@ The deterministic Landau-Lifshitz-Gilbert equation conserves the moment magnitude


anyway. However, as we shall see below, this is not automatically the case for its stochastic
analogue Eq. (10.40) .
474 D. V. Berkov, N. L. Gorn

known (K loeden, 1995) that if one adds to the system of stochastic ODEs
X
dd ; = A;(x,t) + ~B;k~k (10.47)
t k

the deterministic drift term 0 ~ jkB jk (3 B ik j() Xj)' then the Ito solution of this
new system

dx i -_ A;(x,t)
-d +0 '"
LiBjk 3 B ik
--,- + 'LiBik~k
" (10.48)
t jk U Xj k
is equivalent to the Stratonovich solution of the initial system Eq. (10.47).
Comparing Eq. (10. 47) with the LLG-system Eq. (10.46) which we are
interested in, we can immediately see that in our case the matrix B is B ik =
' " 1.E ;jkm j' so that the drift term 0'"
Li Li lk. B jk (3 B;k /3 Xi) reduces to
dm
dT' =-2Dm;.

This drift is directed along the magnetic moment mi' thus trying to change its
magnitude which is forbidden by the model. For this reason this term must be
discarded, which means that for stochastic dynamics of models with rigid
dipoles (dipoles with constant magnitudes) there is no difference between the
Ito and Stratonovich solutions of corresponding stochastic ODEs.
The mathematical reason why the multipl icative noise in Eq. (10.46) does
not lead to the difference between its Ito and Stratonovich interpretations is
that Cartesian coordinates of magnetic moments are not independent
variables: due to the condition that the magnitude of each moment should be
kept constant they are subject to the restriction m~.i + m~.; + m~.; = 1. The
independent variables in this case are spherical coordinates (e, 4 of the
magnetic moment unit vector m. After transition to these coordinates the
stochastic part of Eq. (10. 46) which we have to analyze reads (Brown,
1963b; Braun, 2000)

de = h'l
dT ~,

dcp =_ _l_hfl
dT sine 8

(we have omitted the moment index ; for simplicity), so that the matrix B
responsible for the drift mentioned above is

B = (Boo B~)= ( 0 1
0 ).
B~ B~ - l/sine
It is straightforward to verify that this drift is exactly zero:
D~jkBjk(3Bik/()xj) = o (here ;,j,k=1,2andxl=e, X2=CP) Hence we
arrive at the same result that Stratonovich and Ito stochastic integrals are
equivalent in this case.
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 475

At this point we would like to mention that the opposite statement was
made in Garcia-Palacios and Lazaro (1998) where the authors claimed to
show that Ito and Stratonovich stochastic calculus are not equivalent in
stochastic micromagnetics. The authors of Garcia-Palacios and Lazaro (1998)
have used the formalism of the Fokker-Planck equation (FPE) which describes
the temporal and spatial evolution of the probability distribution of the
magnetization orientation P (m. t). They have demonstrated that an additional
drift term 3(mP)/3m arises in the FPE derived from the Ito interpretation of
the Langevin equation when compared with the Stratonovich one (see p. 14.
940 in Ref. (Garcia-Palacios and Lazaro. 1998. Unfortunately. Garcia-
Palacios et al. did not take into account exactly that point which we have
emphasized in our study: that Cartesian coordinates of the magnetic moment
are not independent. For this reason one cannot use the FPE written in these
coordinates to compare Ito and Stratonovich without introducing the restriction
I m I = 1 explicitly. In particular. the additional drift term 3 (m P) /d m
appearing in the Ito interpretation of FPE should be excluded from this equation
because it leads to the drift of the probability density along the magnetization
vector. This can be clearly seen after transition to spherical coordinates of m.
e.
(m q where this drift term reduces to 3 (mP) /3 m. This means that it tries
to change the moment magnitude. which is again forbidden by the model. We
would like also to add that this mistake does not influence the interesting
physical results obtained in Garcia-Palacios and Lazaro (1998). which
contains a comprehensive study of the single-particle thermodynamics.
To support our conclusion concerning the equivalence of the Ito and
Stratonovich integrals when solving (Eq. (10. 42. we have performed
numerical experiments simulating equilibrium and non-equilibrium properties of
a disordered system of magnetic dipoles. We have solved the stochastic LLG
(Eq. ( 10. 42 using methods converging either to its Ito (Euler scheme) or
Stratonovich (drift-modified Euler and Heun schemes) solutions. We note that
by numerical solution of Eq. (10. 46) Cartesian coordinates are often
preferred. because no instabilities like those observed in spherical coordinates
near the polar axis can occur. During such simulations one has to normalize the
moment vector mj after each new integration step (and also by evaluating the
derivatives at the intermediate points. if necessary) in order to conserve the
moment magnitude.
As the first example we have computed the equilibrium energy distribution
for a single magnetic particle with the uniaxial anisotropy energy
Ean=-KVCos 2 lp only (here V is the particle volume. K-its anisotropy
constant and lp-the angle between the particle moment and its anisotropy
axis). For this purpose we have simulated the motion of a single magnetic
moment without the external field solving (Eq. (10.42 (for a relatively small
damping i\ = O. 1 and a moderate temperature kT / KV = 1. 0) using different
numerical methods mentioned above. After the system has reached the
thermodynamic equilibrium (which may be verified. e. g . by checking that the
476 D. V. Berkov, N. L. Gorn

energy does not exhibit any systematic change) we have started to record the
particle energy at each integration step. After a sufficiently long simulation
time the distribution of these energy values must coincide (in frames of
statistical errors) with the corresponding equilibrium Boltzmann distribution.
The latter can be easily calculated analytically for such a simple system and is

P ( E) cc exp (- E! n _1_ (10.49)


FE
where the reduced energy E = - O. 5cos2 lIJ may vary in the interval [ - O. 5, OJ.
The inverse square root in Eq. ( 10. 49) comes from the density of states in
spherical coordinates.
Energy distribution histograms obtained from the simulations employing
different integration schemes are shown in Fig. 10.20. Here numerical results
are compared with the analytical distribution (Eq. ( 10. 49 displayed as the
thin solid line (the region near the zero energy value is not shown because of
the inverse square root singularity in Eq. (10.49) ). It can be clearly seen that
all three histograms-obtained with CD the Euler method (Ito solution),
CZ) Euler method augmented with the drift term from Eq. ( 10.48) and @ Heun
method (both Stratonovich solutions) perfectly coincide with the analytical
result.
3.0 3.0 3.0

~2.0 ~2.0 ~2.0


c:.. c:.. c:..
1.5 1.5 1.5

1.0 '-------'--_-'-----'-_--'--------J I. O'-------'--_-'-----'-_--'--------J 1.0'-------'--_..1......------'-_---'--------.1


---D.5 ---D.4 -0.3 ---D.2 ---D. I 0 ---D.5 -0.4 ---D.3 ---D.2 -0.1 0 ---D.5 ---D.4 ---D.3 -0.2 -0.1 0
& & &

W ~ ~

Figure 10. 20 The equilibrium energy densities for a single particle magnetic moment
(with the uniaxial anisotropy, temperature kT / KV = 1. 0, damping A = O. 1) obtained by
solving Eq. (10.42) with different numerical methods converging to Ito (a) or Stratonovich
(b), (c) solutions. The equivalence of both approaches can be seen.

The second example deals with magnetic relaxation of a disordered fine


particle system. For this study we have chosen a disordered system of
identical particles with the same uniaxial anisotropy E an = - KVcos 2 lIJ and
aligned anisotropy axes. In such a system without the dipolar interaction all
particles have the same energy barriers 6.E = KV separating two energy
minima along the two opposite directions of the anisotropy axes. Thus,
magnetic relaxation from the state where all magnetic moments are aligned
along one and the same direction of the anisotropy axes should follow the
exponential law m ( t) -exp ( - T / T c), where the relaxation time T c depends
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 477

exponentially on the relation IlE / kT. Corresponding simulation results for


particles with the reduced anisotropy constant f3 = 2 K / M~ = 2. 0 and the
damping A=O.l at the temperature kT/KV=0.5 are presented in Fig. 10.21a.
Here the m ( T )-relaxation curves obtained using the same three numerical
methods listed above are shown. Again, the perfect coincidence (in the frame
of statistical errors) of Ito and Stratonovich solutions can be seen. The inset in
Fig.l0.21a shows the same results in semilogarithmic coordinates to
demonstrate the exponential relaxation of the magnetization expected for a
non-interacting system of identical particles.
Results for the interacting system (particle volume fraction c = O. 08,
initial metastable state prepared by the quasistatic energy minimization
starting from the aligned state) are shown in Fig. 10.21b. In this case the
magnetodipolar interaction between the particles leads to the distribution of
the energy barriers resulting in a non-exponential relaxation. However, the
relaxation curves obtained with different numerical methods perfectly coincide
again, demonstrating the equivalence of Ito and Stratonovich stochastic
calculus for simulation of dynamic remagnetization processes also.

1.0
LOO~
0.10
1.0
- - Euler-Ito
0.8 0.8
" Euler-Strat
:s 0.6 0.01
0 20 40 60 80
:s 0.6 o Heune-Strat

"N 0.4 "N 0.4


0.2 0.2

0 25 50 75 100 0 10 20 30 40
r(r.u.) r(r.u.)
(a) (b)

Figure 10. 21 Magnetic relaxation curves of a system of non-interacting (a) and


interacting (b) particles obtained using the same methods as in Fig. 10. 20. Results
obtained from Ito and Stratonovich stochastic calculus also coincide. (Inset in (a) shows
the same curve in semi logarithmic coordinates) .

Thus we have proved analytically and shown by numerical experiments


that for models with the constant magnitude of the magnetic moment at each
site both Ito and Stratonovich calculus lead to the same physical results despite
that the noise in the stochastic LLG equation is multiplicative. This means that
all results obtained previously using different numerical schemes are correct. A
more important point is that when developing new numerical methods for
solving stochastic equations for such models, one does not need to prove that
these methods converge to the Stratonovich solution; the only consideration
should be the efficiency and accuracy of these new methods.
An alternative numerical method which can be used to study at least the
equilibrium thermodynamical properties of a magnetic system is the well-
478 D. V. Berkav, N. L. Garn

known Monte Carlo method where the system state is updated by performing
either single-moment rotation or collective movements of clusters of magnetic
moments (cluster Monte Carlo methods). The decision whether the system
state obtained after such an update is accepted, is made in a standard way
comparing the energies of the new and updated states (Binder, 1986a).
However, attempts to apply the Monte Carlo method to study dynamic
remagnetization processes encounter two serious difficulties.
First, it is very difficult to establish the relation between one Monte-Carlo
step (a "time" unit in such simulations) and real physical time. Recently,
Nowak et al. Nowak et al. (2000) have suggested the method to calculate this
relation comparing the mean square of the magnetic moment fluctuations near
the equilibrium position of this moment calculated from CD the Langevin
dynamics simulation during the given time interval and CV MC simulations with
the fixed number of MC steps. The vector defining the moment change for each
MC step was chosen randomly from within a sphere of a radius R. The result
of this comparison is the relation between the sphere radius R and the physical
time corresponding to one MC step. However, it is unclear whether the
relation obtained this way also holds for the situation when the moment moves
far away from its equilibrium position, when, e. g., the transition between
metastable states is studied.
Even if this question could be resolved, the second-and more
fundamental-problem of the MC methods would still exist: when comparing
the energies of the old and updated states, dynamical processes which do not
affect these energies cannot be taken into account. The precession of
magnetic moments is such a process, and hence the effect of this precession
can in principle not be included into MC simulations. This means that Monte
Carlo will always remain suitable for the studies of the remagnetization
processes in the high damping limit only.
10.6.1.2 Example 1: ac-Susceptibility for a Disordered Fine
Particle System
Investigation of the dynamical properties (like magnetic viscosity and ac-
susceptibility) of disordered fine magnetic particle systems is one of the most
challenging problems in statistical physics and irreversible thermodynamics.
The long-range and anisotropic dipolar interparticle interaction makes the
development of both analytical and numerical methods for handling these
systems extremely difficult.
As explained above, the only numerical method really suitable for this
task seems to be the Langevin dynamics. In this subsection we present some
our results obtained using this approach for disordered interacting systems of
fine magnetic particles. In particular, we have investigated the temperature
dependence of the ac-susceptibility of such systems for various particle volume
concentrations, single-particle anisotropy values and damping parameters.
We study an interacting system of randomly placed (but non-overlapping)
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 479

identical spherical fine magnetic particles which are assumed to be absolutely


single-domain. This means that the motion of the particle magnetization can be
described as the motion of a single unit vector m j representing the direction of
a "rigid" particle magnetic moment J..l i For this reason the stochastic Landau-
Lifshitz-Gilbert equation for the magnetic moment motion Eq. ( 10.42) may be
used.
In our case the deterministic effective field Hrff includes the external field
xt
Hr , the anisotropy field H~n and the dipolar interaction field H1iP . Below we
assume that all particles possess the uniaxial anisotropy, so that the
anisotropy field is
Han =-2KiVini 0 (mi oni)' (10.50)
The dipolar interaction field can be written as

H. = "" 3e jj (e ij 0 J..l j) - J..l j ,


(10.51)
I L..J 3
i*j 6.r ij
where the usual notation holds. Care must be taken by the evaluation of this
field due to the long-range nature of the dipolar interaction. In principle, for
each particle the sum in Eq. ( 10. 51) must be performed over all other system
particles (for finite systems) or using the Ewald method for systems with
periodic boundary conditions.
We employ periodic boundary conditions to reduce finite size effects, so
that the Ewald method would be the most adequate choice. However, we have
found that for all kinds of simulations presented in this subsection it is sufficient
to use the much simpler extended Lorentz cavity method (Vos et al., 1993;
Berkov, 1996a; Anderson et aI., 1997). In this method the contributions to the
interaction field on the given particle from all the particles inside the sphere
with the "critical radius" R cr around this particle are taken into account exactly
using the summation Eq. ( 10.51). The field from the region outside this sphere
is treated as the field of the homogeneously magnetized media (with the
corresponding average magnetization) inside the spherical cavity. The critical
radius R cr itself should be determined "experimentally" so that its further
increase does not affect the results.
For the solution of Eq. (10.52) we have used the stochastic Heun scheme
(Kloeden and Platen, 1995) which seems to represent the optimal
compromise when trying to achieve the maximal accuracy with the reasonable
programming effort. The time step 6. T used by numerical solution of
Eq. (10.52) was chosen (as usual) as large as possible to perform simulations
over sufficiently long times, but small enough to keep discretization errors in
frames of the statistical accuracy of the results. We have found that the value
6. T = O. 05 was an adequate choice for all parameter sets used in our
simulations.
The real X' (w , T) and imaginary X" (w , T) parts of the ac-susceptibility
were calculated in a standard way. After the thermalization of the system an
480 D. V. Berkov, N. L. Gom

oscillating field
hz = h o cosU> t (10.52)
was applied and real and imaginary susceptibility parts were calculated as

L
1 1 '
X' = -h -L ~ < mi. z > cos ( WT I )
o T 1= 1
L
(10.53)
X" = 1 -L
-h ~< m ,. z > Sin
1 L.J . ( WT, )
o T 1= 1

where < m I.z > is the average z-projection of the system magnetization per
particle at the time step I and T I = I/). T. The reduced frequency is defined as
w = u>y / Ms. Additional checks were performed to verify the linearity of the
system response.
For most results presented below, simulations were carried out on a
system of N p = 256 particles positioned randomly (but non-overlapping) inside
a cubic volume with periodic boundary conditions. For all parameter sets
susceptibility values were sampled during at least N c = 20 oscillating field
cycles. Averaging over at least N coni = 16 (mostly 32) independently generated
spatial particle configurations were performed. Successful tests of our code
were performed on a system of non-interacting particles; results obtained with
our program were in quantitative agreement with known analytical results as
well as with recent numerical calculations performed by Garcia-Palacios et al.
(see Ref. (Garcia-Palacious and Lazaro, 1998) and Ref. therein).
The most interesting question concerning the behavior of the systems
under study is the influence of the dipolar interparticle interaction on the
system dynamics, in particular, on its ac-susceptibility. To study this problem
we have performed simulations for various particle volume concentrations
keeping all other system parameters fixed. Corresponding results are shown in
Figs. 10. 22 and 10. 23, where it can be clearly seen that changes in the
X" ( U> , T) -curves with the increasing particle concentration depend
qualitatively on the single-particle anisotropy value J3 and on the damping i\.
We are interested mainly in the distribution of the free energy barriers in the
systems under study, so that below we shall show results concerning the
imaginary part of the ac-susceptibility only.
High and moderate anisotropy case For moderate and large anisotropies
(J3~1, Fig. 10.22a) the peak on the x"(n-dependencies shifts towards
lower temperatures with the growing particle concentration (increasing
interaction strength). This shift can be clearly seen especially for systems with
low damping: results for J3 = 2. 0 and i\ = O. 1 are shown on Fig. 10. 22a.
When the dissipation increases the peak shift is getting weaker and can hardly
be seen already for the moderate damping i\ = 1. O. These results mean that
the dipolar interaction leads to the decrease of the free energy barriers in fine
particle systems with high and moderate anisotropies, and this effect is more
Numerical Simulation of Quasistatic and Dynamic Remagnetization... 481

0.3 ~ Non-jnt.particles
-----.- c=0.04
---<J- c=0.08
_ _ c=0.16

0.1

o 0.2 0.4 0.6 0.8 1.0


TIEdem
(a)

0.3 -0- w=0.03


----6-- w=O.!
---<>- w=0.3
--:- 0.2
:::l

~
~
0.1

o 0.2 0.4 0.6 0.8 1.0


TIEdem
(b)

Figure 10. 22 X" ( w, T) for large single-particle anisotropy f3 = 2. 0 for (a) various
particle concentrations at the given frequency and (b) for various frequencies for the given
concentration.

pronounced in systems with low damping.


We have also compared the temperature dependencies of the ac-
susceptibil ities for various frequencies keeping the particle concentration
constant. The corresponding result for f3 = 2. 0 is presented in Fig. 10. 22b,
where X" (w , T) -curves for the particle volume fraction c = 0.08 and reduced
frequencies w = 0.03, O. 1 and 0.3 are shown. It can be clearly seen that with
increasing frequency the peak position is shifted towards higher temperatures,
whereby the peak height decreases (see below for comparison with
experimental results).
Low anisotropy case For sufficiently small anisotropy values (how small
depends on the dissipation parameter) the X" ( w, T) -peak shifts towards
higher temperatures when the particle concentration increases-see an
example in Fig. 10. 23a for f3 = O. 5 and i\ = 1. 0). Such a shift can be easily
understood, because the low single-particle anisotropy f3 means that the
interparticle interaction makes the dominant contribution to the effective field
already for moderate particle concentration, thus leading to the increase of the
average energy barrier height when increasing the interaction strength
482 D. V. Berkev, N. L. Gern

(particle concentration) .
The frequency shift of X" ( w, T) for the particle system with the low
anisotropy f3 = O. 5 is shown in Fig. 10. 23b, where corresponding X" -dependencies
for the frequencies w = O. 003, O. Oland O. 03 are displayed. The peak
position moves towards higher temperatures with increasing frequency for this
system also; the physical explanation of this (probably almost universal)
behavior see, e. g., in (Garcia-Palacios and Lazaro, 1998). Note, however,
that the peak height now increases when the frequency increases. This
behavior is qualitatively different from the high anisotropy case, and thus is
very important when comparing our results with experimental data (see
below) .
2.0

{3=2.0 -----0--- Non-inter.


w=0.03 ---+- c=O.OI
2=0.1 --<>-- c=0.02
-T- c=0.04
-----0--- c=0.08
-----+-- c=0.16

0 0.1 0.2 0.3


TIEdeon
(a)

0.4
-----0--- w=0.003
---+- w=0.0 I
0.3
-----0--- w=0. 0 3

0.1

o 0.1 0.2 0.3 0.4


TIEdeon
(b)

Figure 10. 23 The same as in the previous figure for small single-particle anisotropy
{3=O.5.

Comparison with another theoretical results There is a substantial body


of theoretical work done on this subject, but very few reliable results for
interacting dipolar magnetic systems are obtained up to now-mainly due to
the already mentioned long-range and anisotropic nature of the dipolar
interaction.
To our knowledge, all analytical approaches appl ied for the calculation of
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 483

the ac-susceptibility in such systems are based on the energy barrier


approximation (Binder and Young, 1986), where it is assumed that the system
can be considered as a set of independent subsystems, each having its own
relaxation time. The distribution of these relaxation times is due to different
heights of the energy barriers E for corresponding subsystems. If the
distribution of these barriers p ( E) is known, then the calculation of the
corresponding relaxation time distribution f ( Trei) is straightforward-at least in
the Arrhenius approximation, where Trel-exp(E/T). Afterwards, X"(w, T)
can be obtained simply as the convolution of the f ( Trel) with the kernel
WTrel/[ 1 + (WTrel )2J.
Unfortunately, analytical methods for the calculation of the energy barrier
distribution are available only for the weak interaction case-either in the form
of the local field theory (Morup and Trone, 1994; Hansen and Morup, 1998) or
as the pair approximation (Huang and Lu, 1999). Both methods lead to the
conclusion that the dipolar interaction leads on average to the decrease of the
system energy barriers. This seems to support our result that for large single
particle-anisotropy (weak interaction case, see Fig. 10. 23) the X" ( W , T)-
peak shifts towards lower temperatures which means that the energy barrier
distribution moves towards lower barriers. However, theories of this kind are
not really suitable to check numerical or experimental results concerning the
irreversible processes in magnetic systems, because: CD the effect predicted
by such theories is weak by the very definition of the perturbation approach;
~ the temperature dependence of the energy barrier heights and the features
of the energy landscape near the saddle point are ignored; and @ magnetic
moment precession (which may be important, see Refs. (Garcia-Palacios and
Lazaro, 1998; Berkov et aI., 2002 is not taken into account.
The distribution of the energy barriers can be evaluated rigorously, i. e.,
beyond the perturbation approach, using numerical minimization of the
thermodynamical action for the transition between the two given metastable
states. With this method we were able to show (Berkov, 1998) (see also
below) that in systems of magnetic particles with high and moderate
anisotropies the energy barrier distribution indeed moves towards lower
barriers with increasing particle concentration. However, it is still quite
difficult to draw conclusions concerning the ac-susceptibility basing on this fact
only due to reasons ~ and @ mentioned above.
The ac-susceptibility itself for interacting fine magnetic particles was
studied, to our knowledge, only in Anderson et al. (1997) using the Monte
Carlo method. Bearing in mind that this method is not able to take into account
the precession of magnetic moments (see above), we can compare results
from Anderson et al. (1997) with our data obtained in the moderate damping
regime, where the precession is expected to playa minor role. Simulations in
Anderson et al. (1997) were carried out for particles with the reduced
anisotropy constant 13 = 2 K / M~ ~ 1. 8, so that results of Anderson et al.
( 1997) should correspond to our data for 13 = 2. O. Indeed, the shift direction
484 D. V. Berkov, N. L. Gorn

of the X" (w , n -peak with increasing particle concentration found in Anderson


et a!. (1997) (see Fig. 2 in Anderson et a!. (1997 coincides with our
results. However, the relative peak heights for various concentrations are
considerably different. The same applies to the frequency shift of the
X"(w,n-curves (see Fig. 3 in Anderson et al. (1997: the X"(n-peak
moves towards higher temperatures with increasing frequency as for our
results, but the peak heights increase with w in Anderson et al. (1997),
whereby according to our simulations they should decrease with growing
frequency for f3~2 (Fig. 10. 22b). The most probable reason for this
discrepancy is that the MC approach of Anderson et al. (1997) takes into
account single-particle flips only, thus neglecting collective remagnetization
processes. Such processes are not expected to be dominant in systems with
moderate and high anisotropy, but for moderate anisotropy values they can
nevertheless affect the results quantitatively.
Comparison with experimental data The ac-susceptibility of magnetic
nanocomposites is one of their most important features, both from a
fundamental point of view and by evaluation of their suitability for numerous
technical applications. For this reason it has been studied in a huge number of
experimental papers. From this pool of data we have chosen mainly results
obtained on frozen ferrofluids, because CD these systems can be characterized
reasonably well and ~ very accurate and reliable results are available.
Corresponding measurements are usually performed on systems consisting
of magnetite or maghemite particles. In most studies (Jonsson et a!., 1995;
Zhang et a!., 1996; Jonsson et al., 1998; Djurberg et al., 1997; Taketomi,
1998) the following system behavior was observed:
( 1) Concentration dependence: the X" ( T) -peak moves towards higher
temperatures with increasing concentration c.
(2) Frequency dependence: the X" ( T) -peak moves towards higher
temperatures with increasing frequency w, whereby the height of this peak
increases with growing w.
Such behavior agrees very well (at least qualitatively) with our results for
systems with low anisotropies f3 = O. 2 - O. 5 (Fig. 10. 23). However, there
exists an important disagreement between our results and the experiments
cited above: the anisotropy f3 = 2K/M; (calculated from the K-and M s -
values reported, e. g., in Jonsson et al. (1995, 1998 for the particles
studied is about f3 ~ 1. 5 - 2. O. Hence the opposite shift of the X" ( n -peak
with the particle concentration should be observed according to our simulations
(Fig.10.23a).
The most probable reason for this discrepancy (leaving apart trivial
explanations like the particle aggregation which is quite unlikely due to a very
small particle size) is the following. The value of the anisotropy constant K
reported for the given sample is the average value obtained from the
corresponding low-temperature hysteresis loop (Jonsson et al., 1997).
Different particles contribute to this loop-and hence to the average K value-
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 485

with the weight proportional to their moment J.J. By the susceptibility


measurements, however, particle contributions are proportional to the square
of their moment J.J2 (see, e. g, Ref. (K ittel, 1953)), which means that larger
particles contribute much more significantly to the measured susceptibility,
than to the hysteresis.
On the other hand, it is well-known that the total effective anisotropy of
fine particles, being the sum of the crystallographic, surface and shape
anisotropies, depends strongly on the particle size. For all experimentally
studied systems (Gazeau et a!., 1998; Respaud et a!., 1998; Respaud et a!.,
1999) the total anisotropy was found to decrease with increasing particle size
(probably because the surface anisotropy decreases proportionally to the
decreasing fraction of surface atoms when the particle grows, and the shape
anisotropy decreases because the shape of larger particles is closer to the
spheres, at least when particles remain very small). Hence, larger particles
with larger magnetic moments (which contribute significantly more to the ac-
susceptibility than to the magnetization) have much smaller values of the
effective anisotropy 13. This leads to the shift of the X" ( T )-peaks towards
higher temperatures with increasing particle volume fraction, as it should be
for the low anisotropy case (Fig. 10.22). This effect may be enhanced further
by the increase of the particle saturation magnetization M s with the growing
particle size (Morales et al., 1999).
Results of our Langevin dynamic simulations demonstrate two important
features of magnetic nanocomposites. First, their ac-susceptibility peak can
move with the increasing particle concentration (interaction strength) both
towards lower and higher temperatures, which means that the dipolar
interaction can either decrease the (free) energy barriers in such system-if
these barriers are due a substantial single-particle anisotropy, or increase
them-if they are mainly due to this interaction itself. Second, we point out
that our method takes the dipolar interaction fully into account, incorporating
all the dynamic and static correlations between the particle magnetic
moments. For this reason the qualitative discrepancy between our results and
real experiments on ferrofluids with apparently the same single-particle
parameters shows that some additional features of the fine particle systems
should be taken into account to understand their dynamical properties
adequately. We propose that such a feature could be the size dependence of
the total single-particle anisotropy and the particle saturation magnetization.
10.6. 1. 3 Example 2: Standing Spin Waves in Nanodots
Another problem suitable for studies using the Langevin dynamics technique is
the frequency spectrum and the spatial distribution of magnetic excitations
(standing spin waves) in a thin film element with lateral sizes in the sub-lJm
range. The knowledge of this excitation spectrum is important for various
applications making use of such magnetic structures like MRAM-cells (fast
switching processes) and AMR and GMR sensors (signal-to-noise ratio).
486 D. V. Berkov, N. L. Gorn

As an example we present here results of studying a standing spin wave in


=
a rectangular Permalloy element with the lateral sizes@a x b 1.0 IJm Xl. 75 IJm
and the thickness d = 30 nm discretized in the lateral plane in 80 x 140 finite
elements. The standard Py magnetic parameters (M s = 103 G, A = 10- 6 erg/ em)
were assumed. The cubic anisotropy of crystallites with 20 randomly
distributed anisotropy axes was chosen with the anisotropy constant K cub =
4 x 10 4 erg/ cm 3 . The average crystall ite size was taken to be (D) 40 nm, =
whereby it should be noted that the variation of neither (D) nor of K cub ( in
reasonable limits) affected the results qualitatively. The external field parallel
to the long side of the element (chosen as the Oz-axis) Hz = 600 Oe was
applied. Simulations were performed for the room temperature.
First the static equilibrium magnetization structure was calculated
(equilibrium state for T = 0) using the standard quasistatic micromagnetic
minimization procedure. The resulting configuration is shown as the m x -grey-
scale map in Fig. 10.24, where the well-known state with two closure quasi-
domains on the top and bottom sides of the element can be recognized. This
state was used as the initial state for the Langevin dynamics simulations which
were performed using Eq. (10. 42); the damping constant was set to
=
II 0.01, which corresponds approximately to the damping value obtained on
these samples from the ferromagnetic resonance measurements.

Figure 10.24 The equilibrium magnetization configurations (m x grey-scale map).

It has turned out that the corresponding system of stochastic equations


was extremely stiff (probably due to the presence of the exchange interaction
in addition to the magnetodipolar one). For this reason the standard Heun
scheme exploited in the previous example was found to perform extremely

@ The sizes were chosen to match the experimental design of the patterned structures
studied in a group of B. Hillebrands (Univ. Kaiserslautern, Germany) by the quasielastic
Brillouin scattering; when the chapter was in preparation, the experiments were still in
progress, so that no final comparison with the experimental data was possible.
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 487

poorly, requiring time steps less than 10- 4 (in time units tyM s introduced
above). So we were forced to try other methods, and after several attempts
we have developed the optimized stochastic version of the Bulirsch-Stoer
method with the adaptive step-size control. This method has allowed us to
increase the time step at least one order of magnitude (and in typical cases
even more). Details of the method will be given elsewhere.
After the initial "heating phase" (which can be clearly seen in Fig. 10.25
as the initial energy increase) the equilibrium state was reached; the energy
stops to increase and remains constant in frames of temperature fluctuations.
After the equilibrium was reached, we started to record the deviation of each
moment from its equilibrium position for each time step. The Fourier analysis
of these moment trajectories was performed using the Lomb algorithm (Press
et ai., 1992) developed for the analysis of unevenly spaced data. This
algorithm was needed because the time step was adjusted by the program,
and hence, was not constant. Thus, we were able to extract the whole
information concerning the frequency spectrum and space distribution of the
spin excitations in thermal equilibrium.
-2200

-2300

.-.,
=! -2400
~

-2500

-2600'--_ _L...-_----'_ _----l._ _----'-_ _-----'


o 20 40 60 80 100
tyMs

Figure 10.25 Energy time dependence during the Langevin dynamics simulations for the
rectangular magnetic element shown in Fig. 10.24.

The spectrum of m x -osci lIations averaged over the whole sample (i. e.,
over oscillations of all finite element magnetic moments) is shown in
Fig. 10.26 as a thick solid line. The spectrum was additionally averaged over
N con! = 4 independent runs performed for samples with different crystallite
structures. The magnitude of the corresponding statistical error is shown at the
bottom of the same figure with a thin line. Comparison of the magnitudes of the
spectral peaks with those of the statistical errors shows that probably only the
double peak near the point (a) and the somewhat lower peak (b) represent
real spectral features, all other peaks seem to be in frames of statistical
errors, so that the oscillation spectrum has no sharp features for frequencies
488 D. V. Berkov, N. L. Gorn

f?;:; O. 4 .

o 0.2 0.4 0.6 0.8 1.0


fyM s

Figure 10. 26 Power spectrum of the m x -oscillations averaged over the whole sample
(thin line at the bottom gives the magnitude of the statistical error). Spatial distributions of
the oscillation power for frequencies marked with letters a-h are shown in Fig. 10.27.

Spatial structure of the oscillations corresponding to various spectral


frequencies marked in Fig. 10.26 with letters (a) and (h) is shown in
Fig.l0.27. As expected, the softest modes (with the lowest frequencies)
correspond to the oscillations of the region where two domain walls meet each
other (a) and to the oscillations of the domain walls themselves (b).

(a) (b) (c) (d)

(e) (f) (g) (h)

Figure 10. 27 Spatial distributions of the oscillation power for spectral frequencies
marked in the previous figure with corresponding letters.
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 489

Afterwards the mode structure gradually became more and more complex,
whereby the modes started to "propagate" towards the inner area of the
element, where finally a fairly complex structure with one (e), two (f), three
(g) and four (h) half-waves in the direction perpendicular to the applied field
was formed.
As it was mentioned above, the Langevin dynamics simulations of the spin
excitations take into account all the interactions in the system under study, so
that 'the quantitative comparison with the experiment should be possible.
However, in this case one important approximation was made by the
simulations: the thickness of the film d = 30 nm is clearly larger than the
characteristic magnetic length of Py. Hence simple 2D domain walls shown in
Fig. 10.24 may actually possess a complicated internal structure which would
immediately manifest itself in additional oscillation modes. Simulations of a
rectangular magnetic element with the full 3D discretization are necessary to
study this problem.

10.6.2 Slow Remagnetization Dynamics

10.6. 2. 1 Minimizing Thermodynamical Action or How to Handle


Barriers of Arbitrary Heights
Direct simulation of the magnetic moment trajectories-including all
interactions and cooperative dynamic effects-is the biggest advantage of the
Langevin dynamics: one simply mimics the time-dependent system behavior
for the given temperature without any additional and often hardly controllable
assumptions. However, for a system with moderate and high energy barriers
this advantage turns out to be the worst limitation of the method-after all, one
simply mimics the time-dependent system behavior, and hence, the simulation
time necessary to overcome the barrier exponentially growths with the barrier
height-exactly as for real systems, where the Arrhenius-Van't Hoff law holds
(transition probability per unit time to overcome the barrier with the height IlE
is p-exp(-IlE/T)).
Numerical simulations of magnetic systems with such high barriers in
general and the calculation of the transition probability over these barriers in
particular are extremely important from the practical point of view (not to
mention the fundamental problems in the phase transition theory): they are the
only way to predict the long-time stability of the information storage both for
traditional (hard disks) and novel (MRAM) technologies. In this case we are
interested in storage (stability) times of at least several years, so that the
energy barriers should be much higher than the working temperature in order to
make the transition probability p vanishingly small. Fortunately, for IlE T
the task of evaluation p is somewhat simpler because in this limit we do not
need to know all the details of the transition trajectory; it is often sufficient to
490 D. V. Berkov, N. L. Gorn

find the lowest saddle point between the two metastable states of interest. Its
height gives us the corresponding energy barrier b..E between these states,
thus allowing us to estimate the transition probability using the Arrhenius-Van' t
Hoff law mentioned above.
Unfortunately, even the problem of finding such a saddle point cannot
besolved analytically for an interacting system. The reason is that such a
point, being a point in the system coordinate space {x;} where all energy
derivatives aE /a x; are zero (but where neither a maximum nor a minimum of
E should be achieved!) can be found only as a solution of a system of non-
=
linear (in a general case) equations aE /a x; O. We lack general methods for
the solution of such systems, and there exist even arguments that there will
never be any (see, e. g., Ref. (Press et aI., 1992. Reliable (semi)
analytical methods for the saddle point search in magnetic systems are
applicable only for a single particle case (Braun, 1994; Klik and Gunther,
1990) or when the system remagnetization mode is known and is relatively
simple (Braun, 2000).
In this subsection we describe a very general and powerful numerical
method for the evaluation of the energy barrier between the two chosen
metastable states. The method is based on the search for the most probable
(optimal) transition path between these two states by minimizing the
corresponding action derived from the path-integral formulation of the
problem. The required energy barrier can then be calculated as the barrier
along this optimal trajectory.
General idea The underlying idea is several decades old and is due to
Onsager and Machlup (1953). To explain it, let us consider a system of N
classical particles with Cartesian coordinates x; and velocities dx; / d t (i = 1,
"', N) in a viscous fluid. If the interaction energy V (x) of the particles
depends on their coordinates only (x = (x 1 , " ' , X N then the equations of
motion for our system in the presence of thermal fluctuations (Langevin
equations) can be written as

(10.54)

where we have neglected the inertial term for the sake of simplicity and
absorbed the friction constant into the time scaling. As mentioned above,
random Langevin forces ~; assuming to simulate thermal effects can be
considered in most cases as independent random variables with the Gaussian
=
distribution and zero correlation time (~; (0) ~i (t) 205 i/i (t .
Due to these simple correlation properties it is quite evident that the
probability of some particular noise realization {~i (t)}, i = 1, "', N for the
time period [0, tJ is given by (Onsager and Machlup, 1953; Feynman and
Hibbs, 1965; Bray and McKane, 1989)
Numerical Simulation of Quasistatic and Dynamic Remagnetization... 491

If

p[s( 0] = Aex p ( - 46f 2:so I


7dt ). (10.55)

Rewriting the system (Eq. (10.54 as Si (t) = dx;/dt + (J V (x)/(J Xi we


immediately obtain that the probability to observe a given trajectory x ( t) for
the transition between the two states A and B during the time t f (XA (O)-XB (t f
is
If

p[x( t)] = C J[x]ex p [ - 46fo


dt x 2: (ddX/
I
+ (J ~;~) 2],
(10.56)
where J[xCt)] is the Jacobian of the variable transformation x-~.
The total transition probability between (A - B, t f) is then given by the
integral over all trajectories (paths) x( 0 :

-f [
X
8

peA - B, t f ] = C OxCt)J[x] x exp - S(X(Ott)]


40 (10.57>
XA

where the action S (xU is defined as

(10.58)

Evaluation of the path integral (Eq. (10. 57 in a general case is a virtually


unsolvable task for any interacting, many-particle system. But for low
temperatures it is obvious that the main contribution to Eq. (10. 57) comes
from the trajectories close to the optimal trajectory X opt ( t), i. e., to the
trajectory which minimizes the action S (x ( t.
The energy barrier for the
given transition can be found as the barrier along such a trajectory:
t:..E(A-B) = E max (xopt) - EA' The problem we are left with is the
minimization of the functional S (x( 0) .
It is, of course, impossible to perform this minimization analytically for
any realistic model. And even numerical minimization of this thermodynamic
action is highly nontrivial (see below), which is the most probable reason why
the whole method was practically abandoned during the last decades. Here we
do not consider the possibility to minimize the functional Eq. ( 10.58) using the
solution of the corresponding boundary value problem for its Euler-Lagrange
equations (for detai led explanations see (Berkov, 1998) ). Instead, we
approximate the integral (Eq. (10. 58 by a numerical quadrature and then
minimize the function of many variables resulting from this approximation.
Dividing the time interval [0, t.] into K slices with the time step t:..t = t l / K
and approximating (Eq. (10.58 by the simplest quadrature, we obtain
492 D. V. Berkov, N. L. Gorn

SdiSC (x) = LltIji:[Xi.k+I-Xi.k +J.-(dV{Xk+l} +dV{Xk})]2


k=O i=1 I::!.t 2 dXi.k+1 dXi,k
(10.59)

where xi,kis the coordinate of the i-th particle at the time t k =kl::!.t. and Xk =
(x i.k' i = 1, "', N) denotes the set of all particle coordinates for the k-th
slice, Thus for a O-dimensional N-particle system we have to minimize a
function of 0 NK variables.
A simple test example The simplest test of this basic idea is the search
of an optimal trajectory between the two minima for some simple energy
landscape. To perform such a test, we have calculated the optimal trajectory
for a particle moving in the 20 space - (XI' X2 )-plane - with the potential
energy

(10.60)

where x = (x I' X2) and r = (r I' r 2) are vectors in the (X I' X2 ) -plane;
amplitudes U j are positive (negative) for the energy maxima (minima); rj
determines the position; ii j is the width of the j-th maximum (or minimum); J
is the total number of maxima and minima of the potential.
The test result for the simple energy surface with two maxima PI and P 2
and two minima M I and M 2 is shown in Fig, 10,28. The minimization of the
action Eq. (10. 59) for the potential Eq. (10, 60) was performed using K =
128 time slices with Llt = O. 25; the starting trajectory was the straight line
between the minima M I and M 2 . The final trajectory shown in Fig. 10.28 as

/' -- - ........ ~'2


I ... -
I

Figure 10.28 A simple energy landscape used to demonstrate the search for the optimal
transition trajectory between the two energy minima. The white solid line is the "true"
optimal trajectory found by the algorithm, the black dashed line is the false minimum of the
action Eq. ( 10.58) (see text for details) .
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 493

the solid white line clearly passes through a saddle point providing the correct
value of the energy barrier separating the two minima M 1 and M z , thus
demonstrating that the idea works (at least in principle) fairly well.
"False" and "true" optimal trajectories Unfortunately, this simple energy
landscape is suitable mainly for demo purposes. Systems of real interest
possess a large number of degrees of freedom. And the major difficulty arising
from the minimization of the discrete action SdiSC Eq. (10. 59) for these
practically relevant cases is not due to the huge number of variables ....... 0 NK on
which this action depends. This major difficulty is also not due to the often very
"unpleasant" behavior of SdiSC as a function of system coordinates which
requires the application of special minimization met.hods like those described
in Section 10.4.3 (Berkov, 1998). The main problem is the presence of many
undesired local minima of the functional (Eq. ( 10.58)), i. e., the presence of
many trajectories between states A and B in the system coordinate space
which minimizes Eq. (10.58) but does not provide any information about the
corresponds energy barriers.
To explain why this is almost always the case we shall need two facts.
The first fact is that for any path minimizing the action Eq. (10. 58) the
conditions dx i j d t = a v (x) ja x i should be fulfilled; the plus (minus) sign
corresponds to the downhill (uphill) trajectory parts. These conditions mean
that the optimal trajectory goes along the gradient lines of the energy surface.
The second fact we need is that the value of the action S (x) for the
optimal path depends only on the sum of the energy barriers which the path has
to climb over: if the optimal trajectory consists of L pieces where the system
moves uphill and L pieces downhill then the action value is
L

Sex) = 4~t.V, (10.61)


1=1

where t. VI is the energy difference between the end and start points of the I-th
uphill piece. Both these facts for a 10 case are proved, e. g., in Bray and
McKane, (1989); the generalization to a multidimensional case is
straightforward.
According to these considerations, both the solid white line M 1 -M z and
the black dashed line M 1 -P z-M z in Fig. 10.28 deliver local extrema to the
action for the transition M 1 - M z , because both paths proceed along the
gradient lines of the energy surface. It is also easy to show that both extrema
in this situation are local minima of the action; in fact, both paths were
obtained by minimizing the action (Eq. (10.59)) with the potential (Eq. (10.60))
starting from different initial trajectories. Both paths are from the mathematical
point of view local minima of the action for the transition M1-M z . However,
these two path clearly have nothing in common in the physical sense: the first
of them (sol id white line) passes through the correct saddle point, thus
providing the desired information about the energy barrier height (the "true"
optimal trajectory). The second path (black dashed line M 1 -P z -M z ) goes
494 D. V. Berkov, N. L. Gom

via the energy maximum supplying no useful information whatsoever ("false"


optimal trajectory). We clearly need a reliable algorithm to distinguish
between these two kinds of the optimal paths, otherwise the whole method will
be absolutely useless, because the number of "false" optimal trajectories
rapidly grows with the complexity of system.
Such an algorithm can not be based on checking only the action value for
the given trajectory. Such a check might work for simple cases as shown in
Fig. 10.28, where the action value for the "false" optimal path is larger than
for the "true" optimal trajectory: the "false" path in this case has to climb
higher than the "true" one and hence, according to Eq. ( 10. 61 ), the action
value for the "true" path is smaller. But already for a landscape which is a
little bit more complicated, this is no longer the case. If, for example, we are
looking for the optimal trajectory between the minima M, and M2 (Fig. 10.29),
then the "true" optimal path (solid white line in Fig. 10.29) goes through one
of the intermediate minima M 3 or M 4 which are "on the way" from M, to M 2
The "false" optimal trajectory along the gradient lines may be the path M 1 -
P 2 -M 2 (dashed black line). In this case the action value for the "false" path
may be even smaller than for the "true" one, because the "true" path has to
climb over two energy barriers and the "false" trajectory must surmount only
one peak (P 2 ). This example clearly demonstrates that in order to recognize
"true" optimal trajectories, we should be able to distinguish between
"optimal" trajectories passing CD through energy saddle points ("true" optimal
path) and ~ through local energy maxima ("false" case).

Figure 10.29 A more complicated energy landscape. The action Eq. (10.58) along the
"false" optimal path M,-P2-M2 may have the same or even smaller value than for the
"true" optimal path M,-M 3 -M 2 (see text for details).

An apparently straightforward possibility to discriminate between these


two cases is the study of the curvature tensor of the energy surface at the point
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 495

where the energy along the optimal trajectory reaches its maximal value (or at
several points, if there are several energy maxima along the trajectory). If the
corresponding matrix (of the second derivatives of the system energy) has
exactly one negative eigenvalue, then this energy maximum along the
trajectory indeed corresponds to a transition saddle (Kramers et aI., 1990)
and we have a "true" optimal path. Otherwise, we have either an energy
maximum (all eigenvalues negative) or some other kind of an energy
extremum (saddle with more than one negative eigenvalue does not normally
correspond to the point really separating attraction domains of two energy
minima (Kramers et aI., 1990)).
Unfortunately, this method encounters serious technical difficulties: due to
CD the representation of the continuous trajectory as a discrete set of points
(Eq. (10. 59)) and CV the finite accuracy by the determination of the
"optimal" trajectory the energy maximum along this trajectory does not
normally correspond exactly to the energy extremum on the multidimensional
energy surface. This means that we have to organize the search of the true
energy extremum starting from the energy maximum along the trajectory
solving the non-linear set of equations aE/ax; =0 mentioned above. In this
particular case our chances to find such a solution are quite good, because we
start from the point (extremum along the trajectory) which should be quite
close to the solution (extremum on the multidimensional surface). However,
we have observed that the search may still fail-probably due to the very
complex energy landscape. Another problem of this algorithm is that the
search of the eigenvalues is a time consuming (,.". N 3 operations) and delicate
procedure requiring a high accuracy by computing the second derivative
matrix.
For this reason we have also adopted an alternative strategy to check
whether the energy maximum along the optimal trajectory corresponds to a
saddle point or a maximum. Namely, we jump a little bit from the trajectory
point with the highest energy in a random direction and then try to minimize the
system energy starting from this new position. If all energy minima found this
way coincide either with the minimum from where our trajectory started or with
the minimum where it finished then we conclude that the trajectory indeed
passes through saddle points. If any energy minimum found after such random
jumping and subsequent minimization does not coincide with the start or end
points of the trajectory, then we assume that the trajectory does not pass
through a saddle point and do not use this trajectory when calculating energy
barriers.
In the example presented in Fig. 10. 29 this algorithm works as follows.
Starting from any point a little bit aside from the saddle between M 1 and M 3 ,
we would (after minimizing the energy) obviously land in the minima M, or M 3
which are the terminating points of this trajectory. Hence, we would conclude
that it really passes through the saddle point. But when applying this method to
the path M, -+ P z -+ M z and jumping from point P z , we would (after the
496 D. V. Berkov, N. L. Gorn

subsequent energy minimization) finish quite likely in the minimum M 4 which


does not lie on the path under study (M1-P2-M2)' thus concluding that this
trajectory does not pass through a saddle point.
Other important questions related to this discrimination algorithm (how far
apart from the trajectory maximum should we jump before starting the energy
minimization from this new points, how many such attempts should be carried
out to assure that we did not miss any undesired minimum, etc.) are
addressed in Berkov (1998). Here we would like only to mention that both
discriminating strategies outlined above give the same answer when applied to
simple energy landscapes. For systems with the complex energy surface
possessing a large number of metastable states the two methods do not always
provide the same diagnostics for each particular trajectory, but the distribution
densities of the energy barriers calculated using these two methods always
coincide in frames of statistical errors.
Determination of the transition time There exists apparently an
important contradiction in the whole formalism-on one hand, we intend to
calculate the optimal transition trajectory by minimizing the action (Eq. ( 10. 58 ,
then determine the energy barrier using this trajectory and finally evaluate the
transition time between the two minima using, e. g., the Arrhenius law. On the
other hand, the transition time t I is expl icitly present in the action (Eq. (10. 58 as
the upper integral limit. Hence should be known in advance to set the time
step and/or the number of time slices in the discrete version (Eq. ( 10. 59 .
This means, strictly speaking, that for a rigorous determination of t f one
should at least minimize the action also as a function of this time (actually the
maximization of the total probability (Eq. (10.57 as a function of t f is
desirable, but hardly possible, as mentioned above).
Fortunately, we can make use of the fact that the height of the energy
barrier depends on the transition time only relatively weakly (Berkov, 1998).
For this reason we have used the following method for the tl-determination:
we have minimized the action (Eq. ( 10.59 with the small constant time step
(usually !:It = O. 25 - O. 5) and various numbers of time slices K I beginning
from some relatively small number (usually K I = 16) and doubling it (K I + 1 =
2K I) for the next (I + 1)-th action minimization. The process was terminated
when the relative difference between the two values of the energy barriers
obtained for the subsequent action minimizations was less than t:,.E = 10- 2 .
This way we could determine the energy barrier with sufficient accuracy and
estimate the transition time using the Arrhenius law or more sophisticated
approaches (see below).
10.6. 2. 2 Energy Barrier Density in Magnetic Nanocomposites
To apply the method presented above to the systems of interacting magnetic
moments (classical exchange and dipolar spin glasses, systems of fine
ferromagnetic particles, thin magnetic films, nanodots, etc.) we have to start
with the magnetic counterpart to the Newton-Langevin Eq. ( 10. 53), namely,
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 497

with the stochastic Landau-lifshitz-Gilbert equation of motion for magnetic


moments Eq. ( 10 . 42). Neglecting the inertial term in Eq. ( 10. 43) corresponds
to neglecting the precession term in Eq. (10. 42) (large dissipation A 1)
which for a single magnetic moment leads to the equation

~~ =- {m X [m x (hell + hfl)J} =- m. (m. hie!) + hhel (10.62)

where all constants are again absorbed in the time unit, the total field is hie! =
hell+ h fl and the normalization m = 1 was used by the last transformation.
The conservation of the magnetic moment magnitude makes the transition
to the spherical coordinates (e, cp) of the unit vector m unavoidable, because
only the components of the random field hfl that are perpendicular to the current
moment direction should be taken into account (exactly as only the
perpendicular components of the deterministic effective field are important for
the description of the magnetic moment motion without thermal fluctuations).
Transforming all vectors to the new coordinate system with the z' -axis along
the magnetic moment direction and the x' -axis lying in the meridian plane of
the initial spherical coordinate system (so that in this initial system, as usual,
m x = sine cos cp, my = sine sincp, m z = cose), we obtain equations of motion
for the magnetization angles

e = aE
ae
+h L
x

sine. (P =- _1_ aE + h L (10.63)


sine a y

where h~ and h~ are Cartesian components of the fluctuation field in the new
coordinate system. The components of the deterministic effective field hell are
already expressed as corresponding angular derivatives of the moment
energy.
The generalization to a system of N interacting moments is
straightforward. In the resulting system of the equations of motion for the
moment orientation angles (e;, cp;)

e =- aE{O}
J ae; + h
L
I.X

,

sine; (P; = _ _1_ aE{O} + hL


i = l,,N (10.64)
sine; a; I.y ,

the system energy E { O} (where {O} denotes the set of all angles (e;, cp;))
now includes the interparticle interaction energy of any kind (i. e., exchange,
dipolar, RKKY, etc.). This system is fully analogous to Eq. (10. 53) so that
under the same assumptions (that Cartesian components of the Langevin field
are independent random quantities with the Gaussian distribution and 0-
correlated in time), the transition probability between the two chosen
magnetization states 0 A and Os is
498 D. V. Berkov, N. L. Gorn

P[QA - QB' trJ cc J:: OQ(t)J[Q - hL J x exp [ - S(Q~6' tf) l


(10.65)

Thermodynamical action S for a magnetic particle system is defined as

S[Q(t)J = ftfdt
o
~
;
[(d8;
dt
+ JE{Q})2 +
J8;
(sin8; d</>;
dt
+ ~.1~
s1n8;
JE{Q}
J</>;
)2 .
J
(10.66)

The magnetization path in the Q-space which minimiZeS this functional can
provide information about the energy barrier separating the states Q A and QB
exactly in the same way as for the test system discussed in the previous
subsection.
To find this optimal path, we have used the numerical quadrature
representation of Eq. ( 10. 66)

S disc ( Q) = t::.t Ij t[
k=O ;=1
8;, k+ I - 8 i.k
li.t
+ 21 (J E { Q k+
J8;,k+1
I } + JE { Q k } ) J
J8;,k
2

+ [Sin8;'k+' + sin8;,k </>i,HI - </>;,k


2 t::.t
2
+-.l( 1 .JE{Qk+l}+_1_.JE{Qk})J (10.67)
2 sin8;,k+l J</>;,k+l sin8;,k J</>;.k

analogous to Eq. (10.59). Minimizing the corresponding discrete action SdiSC


as a function of orientation angles (8;, k' </>;, k) of all particles for the time
slices k = 1, "', K -1 (moment coordinates for the O-th and K-th time slices
are fixed being the given coordinates of the initial and final states) we obtain
the optimal system trajectory. The check procedures described above should
be applied to each trajectory found this way to ensure that it really passes
through the saddle points (is "true" optimal trajectory). The energy barrier
encountered along such a "true" optimal trajectory is then assumed to be the
lowest energy barrier between the states Q A and QB'
Apart from the usual difficulties encountered by the minimization of the
many variable functions, the minimization procedure for the discrete action
(Eq, (10.67) is subject to the stability problem specific for the spherical
coordinates: the factors 1/ sin8; in Eq, (10, 67) diverge for any trajectory
closely approaching (at least at one slice) the polar axis of the spherical
coordinate system (8-0 or 8 -1f). For this reason we first have to choose
suitable spherical coordinates for each particle separately at the beginning of
the minimization procedure, However, because particle trajectories are
changing during the minimization, the trajectory of some particle may become
too close to the polar axis of its coordinate system even if initially it was not.
Hence, we have also to watch for such" dangerous" cases and switch to
another spherical coordinate system for the corresponding particle when
necessary (see also Section 1O. 4. 1), For these reasons we have minimized
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 499

the thermodynamical action (Eq. ( 10. 67 using the version of the MSDR
relaxation method (see Section 10. 4. 3), which was found to perform very
quickly for this task (S - 10 times faster than CGM) .
A simple test of our code was performed on a single particle with the
uniaxial anisotropy energy E an = KV sin 2 e in the absence of the external field.
The magnetic moment of each particle in this situation has two equivalent
(meta) stable equilibrium states (along the two opposite directions of the
anisotropy axis) separated by the energy barrier t:.E = KV. The energy
barrier found by our algorithm always agreed with this value within the
numerical accuracy.
Results for magnetic nanocomposites and discussion Using this method,
we have calculated the distribution of the energy barriers in the same system
as described in Section 10. 6. 1. 2-magnetic nanocomposite consisting of N
single-domain nanoparticles with the uniaxial anisotropy Eq. (10. SO).
Particles were embedded in a non-magnetic matrix, so that only the
magnetodipolar interaction (Eq. (10. S1 (treated with the extended Lorentz
cavity method) between the particles was present. Periodic boundary
conditions were assumed.
In the absence of the interparticle interaction and the external field such a
system has 2 N metastable states (energy minima) with the same energy.
Among them the two states were chosen arbitrarily and the algorithm had to
find the connected path between these two states passing through the saddle
points only-a very non-trivial task analogous to the test example presented in
Fig. 10.29, but now in the 2N-dimensional space. From the physical point of
view it is evident that transitions between the local minima of this system occur
via the single-particle moment jumps between the opposite directions of the
particle anisotropy axis. In all studied cases the algorithm indeed was able to
find the path consisting of the single-moment flops between the two chosen
metastable states. The energy barriers along this path for the i-th moment
jump again agreed very well with the analytical value t:.E; = K ;V; .
The most intriguing question for this system is, as already mentioned, the
influence of the magnetodipolar interaction on the system properties, in
particular on the distribution density of the energy barriers p (E) (Hansen and
Morup, 1998; Dormann et al., 1999), which controls both the reversible and
irreversible thermodynamics of the system. To solve this question, we have
computed p (E) for various volume concentrations of the magnetic phase, thus
varying the interaction strength.
Calculations were performed, as for the Langevin dynamics simulations,
for two kinds of systems-with high (f3 = 2 K / M~ = 2. 0) and low (f3 = O. 5)
single-particle anisotropies. Typically the distribution of the energy barriers
was accumulated from N con ! = 4 - 8 realizations of the particle disorder; for
each configurations about N tans = 200 transitions between metastable states
were analyzed. Corresponding results are shown in Figs. 10. 30 and 10. 31 ,
were the distribution of the reduced energy barriers E = E/ M~ V (a) together
500 D. V. Berkov, N. L. Gorn

His!. of the energy barriers Hist. of the moment changes

~5lJl c=O.OI

I I
i :[
0 0.25 0.50 0.75 1.00 0 2 3 4
e !J.m

~5~, c=0.04

I I
I:l ~,
0 0.25 0.50 0.75 1.00 0 I 2 3 4
e !J.m

~s[sL
0 0.25 0.50
1

e
c=0.08

0.75
1

1.00
I:l 0 1
!J.m
2 3
='
4

~~
c;:: 5

0
I

0.25 0.50
c=0.16

0.75
!

1.00
I:l 0
,e-- ;=
2 3 4
e !J.m

~{~ 0 0.25 0.50


c=0.24

0.75
!

1.00
I :~
0 2 3 4
e !J.m
(a) (b)

Figure 10. 30 Density of the energy barriers (left) and magnetization changes (right
histogram) corresponding to the transition between randomly chosen energy minimum in a
system of magnetic particles with the low single-particle anisotropy {3 = O. 5. Dashed line
on the left histograms represent the position of energy barrier for a single particle with {3 =
0.5.

with the density of the magnetization changes corresponding to the transition


between metastable states (b) are presented.
First of all, it can be seen that for low particle concentrations (~1 % )
P (E) consists of the relatively narrow peak positioned at the value
corresponding to the energy barrier Esp = {3/2 for a single particle moment flip,
as it should be for a weakly interacting system. The position of this single-
particle flip barrier is shown both in Figs. 10.30 and 10.31 with the dashed
line.
When the concentration increases the energy barrier density is getting
broader, but for the systems with the low and high anisotropy this broadening
occurs in a qual itatively different ways. For the high-anisotropy case (Fig. 10. 31)
Numerical Simulation of Quasistatic and Dynamic Remagnetization . .. 501

Hist. of the energy barriers Hist. of the moment changes

*:f X~'
0 0.5 1.0
e
1.5
I

2.0
(t JL
I 2
!'im
3
I

*:~~
0 0.5 1.0 1.5
I

2.0
~ :t
I
~
2 3
I

e !'im

*:~~O~8
0 0.5 1.0 1.5 2.0
(t I
~
2 3
I

*:t:==L~-D1,6
e !'im

0 0.5 1.0 1.5


I

2.0
~ :1
1
~
2 3
!

e !'im

*:f~r-O'24 !
~:L .3
0 0.5 1.0 1.5 2.0 I 2
e !'im
(a) (b)

Figure 10.31 The same as in Fig. 10. 30 for the high anisotropy case f3 = 2. O.

the broadening of p (E) with the increasing particle concentration is


accompanied by its shift towards lower energy barriers, so that already for
moderate particle concentration (~4 %) almost all barriers lie below the
E-value for a single particle.
For the system of particles with low anisotropy (Fig. 10.30) barriers
which are both higher and lower, than for a single particle, arise. The
average energy value for the resulting spectrum of energy barriers for this
particular anisotropy still exhibits a minor shift towards lower energy values.
However, there clearly exists a heavy tail of high energy barriers, which is
especially pronounced for concentrations c ~ 12 %. Computing the same
distribution for the system with the lower anisotropy still (f3 = O. 2) we have
observed the overall spectrum shift towards higher energies with increasing
particle concentration.
The distribution density of the energy barriers p ( E), having highly
502 D. V. Berkov, N. L. Gorn

interesting system characteristics from the fundamental point of view, cannot


be compared directly with the experimental data. Hence, the establishing of
the relation between p (E) and measurable quantities like ac-susceptibility and
magnetic viscosity is highly desirable. This is by no means a trivial task at
least for the following reasons (apart from the obvious one that real systems
always have some distribution of single-particle parameters which should be
known to allow for quantitative comparison) : CD all experiments are performed
at finite (and usually even not at low) temperatures so that the density of free
energy barriers is required for their interpretation, ~ different moment
changes occur by transitions over different barriers and @ due to the
interparticle interaction each transition can, in principle, change the height of
other barriers.
The first problem mentioned above is the most difficult one, because to
calculate the height of free energy barriers, we at least have to take into
account the energy landscape near the saddle point. This can be done in
principle analytically by studying the curvature matrix of the energy surface,
but here we encounter the same problems as mentioned in the discussion of the
criterion for distinguishing between "true" and "false" optimal trajectories.
Another way would be to calculate relative probabilities of the trajectories that
are close enough to the optimal one. This would be an attempt to calculate
numerically the path integral in the saddle point approximation (Eq. (10. 17 and
we expect it to be very time-consuming.
The second problem-that different transitions cause different moment
changes-can be solved much more simply. Namely, we only have to keep
record of the differences between the magnetizations /::;.mAB = mA - mB of the
two states A and B for each transition studied. The corresponding histograms
of the /::;.m-distributions p (/::;'m) are shown on the right-hand side in Figs. 10. 30
and 10.31. It can be seen that for weakly interacting systems the
corresponding distributions exhibit a sharp peak near m = 2. This is simply due
to the fact that the moment change corresponding to a single particle moment
flip m--m is I/::;.ml = Im- (-m) 1= 12ml =2 (recall that m denotes the unit
moment vector). With the increasing particle concentration the distribution
p(/::;.m) broadens, signaling the appearance of collective remagnetization
processes.
The key question by the analysis of the moment changes is whether the
magnitude of these changes is correlated with the height of the corresponding
energy barrier. If, e. g., it would turn out that the moment changes tend to
zero when the energy barrier height for the transition decreases, this would
mean that the small energy barriers do not play any significant role in the
system thermodynamics, because corresponding magnetization changes are
hardly noticeable. We have shown that is not the case just by plotting the 20
mutual distribution of the energy barriers and moment changes p ( E , /::;.m ) .
Corresponding grey-scale plots for a system with low anisotropy f3 = O. 5 and
two different concentrations are shown in Fig. 10. 32. The picture for the
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 503

2.5 2.5

::: 2.0 2.0


<i

1.5 1.5

1.0 1.0
0.1 0.2 0.3 0.1 0.2 0.3
tlE M.
(a) (b)

Figure 10. 32 Mutual 2D distribution density p ( E, I:>.m) of the energy barriers and
moment changes for di lute (part (a), c = O. 01) and concentrated (part (b), c = O. 24 )
magnetic nanocomposites with f3 = O. 5.

lowest concentration c = O. 01 is, as expected, very simple: p ( E, f::,.m )


consists of a single sharp peak positioned at the point (E = 0.25, f::,.m = 2.0) .
From the density p (E , f::,.m) for the high concentration c = O. 16 it can be seen
that, although the moment changes for the low barriers are concentrated at
somewhat smaller values than f::,.m for the higher ones, they do not tend to zero
at all. Hence all transitions provide approximately equivalent contributions to
the system thermodynamics.
A preliminary discussion concerning the comparison with the available
experimental results may be found in Berkov (1998) and in Section 10. 6. 1. 2.
Here we would like only to mention that the qualitative difference in the
behavior of the energy barriers distributions with increasing particle
concentration (interaction strength) for systems with low and high single
particle anisotropies obtained using the path integral method (Figs. 10.30 and
10. 31) seem to support results obtained using the Langevin dynamics
formalism for the ac-susceptibility of these systems (see Section 10.6. 1.2).

References
Abramovitz, M., I. Stegun. Handbook of Mathematical Functions. Dover
Pub!., New York (1968)
Acton, F . S. Numerical methods that (usually) work. Math. Ass. of America,
Washington D. C. (1990)
Altbir, D., P. Vargas, J. d' Albuquerque e Castro, U. Raff. Phys. Rev. B 57 :
13,604 (1998)
O. Anderson, J., C. Djurberg, T. Jonsson, T. Svendlindh, P. Nordblad. Phys.
Rev. B 56 13,983 (1997)
504 D. V. Berkov, N. L. Gorn

Advanced Recording Model. package for micromagnetic simulations. Euxine


technologies. http://www.euxine.com
Asselin,P., A. Thiele. IEEE Trans. Magn. MAG-22: 1876 (1986)
Berkov,D.V., S.V.Meshkov. Sov.PhysJETP. 94:140 (1988)
Berkov, D. V., S. V. Meshkov. IEEE Trans. Magn. MAG-26: 1804 (1990)
Berkov, D. V., K. Ramsteck, A. Hubert. Phys. stat. sol. 137: 207 - 225
(1993)
Berkov,D.V., J. Magn. Magn. Mat. 161: 337 (1996a)
Berkov,D. V., N. L. Gorn, W. Andra, R. Mattheis, S. Schmidt. Phys. stat.
sol. (a) 158: 247-258(1996)
Berkov,D.V. Phys. Rev. B 53:731-734 (1996b)
Berkov,D. V. J. Magn. Magn. Mat. 186: 199 (1998)
Berkov,D.V., N.L.Gorn. Phys. Rev. B57: 14,332,343 (1998)
Berkov, D. V., N. Gorn, P. Gernert. Numerical simulations of remagnetization
processes in sub-mkm multilayer magnetic structures. In: S. Mengel ed.
Proceedings of the Statusseminar "Magnetoelektronik", p. 95(2000)
Berkov, D. V., N. L. Gorn, P. Gernert. J. Magn. Magn. Mat. 226: 1936
(2002)
Bertram, H., H. Schabes. J. Appl. Phys. 64: 1347 (1988)
Bertram, H. N., C. Seberino. J. Magn. Magn. Mat. 193: 388 - 394 (1999)
Binder, K. Monte-Carlo Methods in Statisitical Physics. Springer-Verlag,
Berlin, 1986
Binder,K. A.P.Young. Rev. Mod. Phys. 58: 801-976 (1986)
Braun, H.-B., J. Appl. Phys. 76: 6310 (1994)
Braun, H. -B. Stochastic magnetization dynamics in magnetic nanostructures:
from Neel-Brown to soliton-antisoliton creation. In: P. Entel & D. Wolf,
Eds. Structure and Dynamics of Heterogeneous Systems. World Scientific
(2000)
Bray,A. J. Jr. A. J. McKane. Phys. Rev. Lett. 62:493 (1989)
Brown, W. F., Jr. Micromagnetics. Wiley, N. Y. (1963a)
Brown,W. F., Jr. Phys. Rev. 130: 1677 (1963b)
Brown, W. F. A. E. LaBonte. J. Appl. Phys. 36: 1380 -1386 (1965)
Chikazumi, S. Physics of Ferromagnetism. Clarendon Press, 1997
Chui,S. T. J. Appl. Phys. 79: 4951 (1996)
De' Bell, K ., A. B. Macisaac, J. P. Whitehead. Rev. Mod. Phys. 72: 225-257
(2000)
Deserno,M., C.Holm. J. Chem. Phys. 109: 7678(1998)
Djurberg, C., P. Svendlindh, P. Nordblad, M. F. Hansen, F. Bodker, S. Morup.
Phys. Rev. Lett. 79: 5154 (1997)
Dormann,J.L., D.Fiorani, E.Tronc. J. Magn. Magn. Mat. 202: 251(1999)
Dotsenko, V. S. Sov. Physics-Usp. (USA). 36: 455 (1993)
Feng, X., P. B. Fischer. J. Appl. Phys. 80: 6988 (200 1)
Feynman, R. P., A. R. Hibbs. Quantum Mechanics and Path Integrals.
McGraw-Hili, N. Y. (1965)
Numerical Simulation of Quasistatic and Dynamic Remagnetization. .. 505

R. Fischer, T. Schrefl, H. Kronmuller, J. Fidler. J. Magn. Magn. Mat. 150:


329 - 344( 1995)
Fredkin, D. R., T. R., Koehler. J. Appl. Phys. 67: 5544 - 5548 (1990)
Fredkin,D. R., W. Chen, T. R. Koehler. IEEE Trans. Magn. MAG-28: 2880
(1992)
Garcia-Palacios,J. L., F. J. Lazaro. Phys. Rev. B 58: 14, 937( 1998)
Gardiner, S., Handbook on Stochastic Processes. Springer-Verlag, Berlin
( 1997)
Gazeau, F., J. C. Bacri, F. Gendron, R. Perzynski, Y. L. Raikher, V. I.
Stepanov, E.Dubois. J. Magn. Magn. Mat. 186:175 (1998)
Gear, C. W. Numerical Initial Value Problems in Ordinary Differential
Equations. Englewood Cliffs, Prentice-Hall, N.Y. (1971)
Gibbons,M.R. J. Magn. Magn. Mat. 186: 389-401 (1998)
Giles, R. C., P. R. Kotiuga, M. Mansuripur. IEEE Trans. on Magn. MAG-27:
3815 (1991)
Gill, P. E., W. Murray, M. H. Wright. Practical Optimization. Academic
Press, San Diego (1981)
Hadjipanayis,G. C. J. Magn. Magn. Mat. 200 :372 (1999)
Hansen,M. F., S. Morup. J. Magn. Magn. Mat. 184: 262 (1998)
Hoenggi, P., P. Talkner, M. Borkoves. Reaction-rate theory: fifty years
after Kramers. Rev. Mod. Phys. 62: 251-341 (1990)
Huang,H.L., J.J.Lu. Appl. Phys. Lett. 75:710 (1999)
Hubert, A., R. Schafer. Magnetic domains: the analysis of magnetic
microstructures. Springer-Verlag, 1998
Hubert, A., W.Rave. Phys. stat. sol. (b) 211 (1999) 815-829 (1999)
Hucht,A., A.Moscher, K.D.Usadel. J. Magn. Magn. Mat. 148: 32 (1995)
Jonsson, T., J. Mattson, C. Djurberg, F. A. Khan, P. Nordblad, P. Svendlindh.
Phys. Rev. Lett. 75: 4138 (1995)
Jonsson, T ., J. Mattson, P. Nordblad, P. Svendl indh. J. Magn. Magn. Mat.
168: 269 (1997)
Jonsson,T., P.Nordblad, P.Svendlindh. Phys. Rev. B 57:497 (1998)
Kittel,C. Introduction to Solid State Physics. J. Wiley, N. Y. (1953)
Klik,l., L.Gunther. J. Appl. Phys. 67: 4505 (1990)
Koch R.H. et al. Phys. Rev. Lett. 81: 4512 (1998)
Kloeden, P. E., E. Platen. Numerical Solution of Stochastic Differential
Equations. Springer-Verlag, Berlin (1995)
Koehler, T. R., D. R. Fredkin. IEEE Trans. Magn. MAG-28: 1239 - 1244
(1992)
Hoenggi, P., P. Talkner, M. Borkovec. Rev. Mod. Phys. 62: 251 - 341 ( 1990)
Kuzminski, M., A. Slawska-Waniewska, H. K. Lachowicz. IEEE Trans. Magn.
MAG-35: 2853 ( 1999)
Labrune, M., J. Miltat. IEEE Trans. Magn. MAG-26: 152l( 1990)
Landau, L. D., E. M. Lifshitz. The Classical Theory of Fields. Pergamon
Press, Oxford (1975)
506 D. V. Berkov, N. L. Gorn

Landau, L. D., E. M. lifshits. Electrodynamics of Continuous Media. Elsevier


Science (1985)
Lopez-Diaz,L., J. Eicke, E. Della Torre. IEEE Trans. Magn. MAG-35: 1207-
1210(1999)
Lyberatos, A.,R.W.Chantrell. J.AppI.Phys. 73: 6501 (1993)
Ramstock, K. FEM software package for micromagnetic simulations. http: / /
www.ramstock.de
Mansuripur,M., R.Giles. IEEE Trans. Magn. MAG-24: 2326 (1988)
Mansuripur, M., R. C. Giles. Computers in Physics. 4: 291 (1990)
McKane,A. J., H. C. Luckock, A. J. Bray. Phys. Rev. A 41: 644 (1990)
McShane, E. J. Stochastic calculus and stochastic models. Acad. Press, New
York (1974)
Miles,J.J., B.K.Middleton. J. Magn. Magn. Mat. 95: 99-108(1991)
Miltat,J., G. Albuquerque, A. Thiaville. An Introduction to Micromagnetics in
the Dynamical Regime. In: B. Hillebrands, K. Ounadjela, Eds. Springer-
Verlag Berlin, (2001) Spin Dynamics in Confined Magnetic Structures I,
Topics in Applied Physics, Vol. 83
Miltat J, G. Albuquerque, A. Thiaville. Micromagnetics: Dynamical Aspects.
In: E. Beaurepaire, F. Scheurer, G. Krill and J. -Po Kappler, Eds.
Magnetism and Synchrotron Radiation, Lecture Notes in Physics, Springer-
Verlag, Berlin (2001)
Morales,M. P., M. Andres-Verges, S. Veintemillas-Verdaguer, M.I. Montero,
C. J. Serna. J. Magn. Magn. Mat. 203: 146 (1999)
Morup,S., E.Trone. Phys. Rev. Lett. 72:3278 (1994)
Nakatani, Y., Y. Uesaka. Jpn. J. Appl. Phys 1, 28: 2485 (1989)
Nakatani, Y., Y. Uesaka, N. Hayashi, H. Fukushima. J. Magn. Magn. Mat.
168: 347 (1997)
Nowak, U., R.W.Chantrell, E.C.Kennedy. Phys. Rev. Lett. 84: 163-166
(2000)
O'Grady,K., H.Laider. J. Magn. Magn. Mat. 200: 616(1999)
Onsager, L., S. Machlup. Phys. Rev. 91: 1505 (1953)
Parker,G.J., C.Cerjan, D.W.Hewett. J. Magn. Magn. Mat. 214: 130-138
(2000)
Press, W. H., S. A. Teukolsky, W. T. Vettrling, B. P. Flannery. Numerical
Recipes in Fortran: the Art of Scientific Computing. Cambridge University
Press (1992)
Ramstock, K., T. Leibl, A. Hubert. J. Magn. Magn. Mat. 135: 97 - 110
(1994)
Ramstock, K. Mikromagnetische Rechnungen an isolierten und eingebetteten
Strukturen, PhD Thesis (in German), Univ. of Erlangen, Germany( 1997)
Respaud, M. et al. Phys. Rev. B 57: 2925 ( 1998)
Respaud,M., M. Goiran, J. Broto, F. H. Yang, T. Ould Ely, C. Amiens, B.
Chaudret. Phys. Rev. B 59: R3934(1999)
Rumelin,W. SIAM J. Numer. Anal. 19: 604(1982)
Scheinfein, R. Landau-lifshitz-Gilbert Micromagnetics Simulator. http: / /
Numerical Simulation of Quasistatic and Dynamic Remagnetization. . . 507

www.dancris.com/llg
Scholz,W., T.Schrefl, J.Fidler. J. Magn. Magn. Mat. 233: 296(2001)
Schrefl, T., J. Fidler, H. Kronmuller. J. Magn. Magn. Mat. 138: 15 - 30
(1994a)
Schrefl,T., J.Fidler, H.KronmUller. Phys. Rev. B 49: 6100-6110(1994b)
Schrefl,T., J.Fidler, J.N.Chapman. J. Phys. D.: Appl. Phys. 29: 2352-
2361( 1996)
Schrefl,T., J. Fidler, K. J. Kirk, J. N. Chapman. IEEE Trans. Magn. MAG-
33: 4182 - 4184 ( 1997)
C. Slater, J. Insulators. Semiconductors and Metals. McGraw-Hili, N. Y.
(1967)
Suhl, H. IEEE Trans. Magn. MAG-34: 1834 ( 1998)
Taketomi,S. Phys. Rev. B 55: 3073(1997)
Tehrani,S., J.M.Slaughter, E.Chen, M.Durlam, J.Chi, M.DeHerrera. IEEE
Trans. Magn. MAG-35: 2814 - 2819( 1999)
Thiaville,A., J. Miltat, J. Ben Youssef. Europ. Phys. J. B 23:37 (2001)
Vos,M.J., R. L. Brott, J. Zhu, L. W. Carlson. IEEE Trans. Magn. MAG-29:
3652( 1993)
Wang,Z.C., D.J.Mapps, L.N.He, W.W.Clegg, D.T.Wilton, P.Robinson,
Y.Nakamura. IEEE Trans. Magn. MAG-33: 4498-4512 (1997)
Williams,W., D.J.Dunlop. Phys. Earth Planetary Interior. 65: 1 (1990)
Wohlfarth, E. P., E. C. Stoner. Phil. Trans. Roy. Soc. (London), Ser. A 240:
599( 1948)
Wright, T. M., W. Williams, D. J. Dunlop. J. Geophys. Res. 102: 12,085-
12, 094( 1997)
Yang,B., D.R.Fredkin. IEEE Trans. Magn. MAG-34: 3842-3851(1998)
Yuan,S. W., H. Bertram. Phys. Rev. B 44 12,395 (1991)
Yuan,S.W., H.N.Bertram. IEEE Trans. Magn. MAG-28: 2031(1992)
Zhang,J., C.Boyd, W.Luo. Phys. Rev. Lett. 77: 390(1996)
Zhang,K., D. R. Fredkin. Thermal relaxation of clusters of interacting Stoner-
Woholfarth particles. J. Appl. Phys. 87: 4795 (2000)
Zhang, K., D. R. Fredkin. J. Appl. Phys. 87: 4795 (2000)
Zhao,Y., H.N.Bertram. J. Magn. Magn. Mat. 114: 329(1992)
Zhu,J.-G. H.N.Bertram. J. Appl. Phys. 63: 3248-3253 (1988)
Ziman, J. M. Models of disorder. Cambridge Univ. Press (1979)

The authors are greatly indebted to Prof. A. Hubert who was their first teacher in
micromagnetics and who continued to support them during the following years.
We also would like to thank Prof. W. Andra for his permanent willingness to share with
us his invaluable experience in virtually all problems of magnetism, Prof. P. Gornert for his
continuous interest and support of our work, Dr. R. Mattheis for providing many interesting
problems to solve and his trust that we are able to solve them, Dr. K. Ramstock for many
fruitful discussions during our stay in Erlangen. We are grateful also to Dr. S. Demokritov,
Prof. B. Hillebrands, Prof. J. Millat, Dr. T. Schrefl and Dr. J. Wecker for their interest
and support of various parts of the research described in this contribution.
11 Preisach Model and Simulation of Relaxation
Kinetics

K. L. Metlov

11. 1 Introduction

The hysteresis is a complicated phenomenon, taking place in non-linear, non-


equilibrium systems. During the evolution under the influence of slowly applied
external force, such systems sequentially "visit" one of many metastable
states. Multiplicity of these states (connected to the essential non-linearity of
the potential) and inability of the system to completely relax during the
evolution (so that it is non-equilibrium at any time) are basic requirements for
systems to have hysteresis. Both these factors result in significant
complications of analysis of such systems.
In this chapter the quasistatic scalar hysteresis (that is, both response
and external force are scalars) is studied theoretically with the framework of
the Preisach model. This model allows representing the evolution of a large
variety of systems, and will be reviewed at the beginning of the chapter. The
conditions of the applicability of the Preisach model based on both the
measurement of the hysteresis data and on the statistical properties of the
potential of the systems will be discussed. The theoretical representations of
simple model systems will be built and the connection of these Preisach
representations to the kinetics of the thermal relaxation of these systems will
be considered. While discussing these phenomena we shall have magnetic
hysteresis in mind, but most, if not all, of the results can be translated to other
non-linear, non-equilibrium systems as well. In magnetic systems under
consideration the macroscopic quantity of interest is the average
magnetization, and the macroscopic external force is the magnetic field. The
scalar case considered here corresponds to the study of the dependence of the
volume averaged magnetization projection on a given direction on the
projection of the magnetic field vector.
One of the important features of the magnetic systems is a large number of
internal degrees of freedom (which can be often considered as being infinite) .
For example, a macroscopic soft magnet is usually subdivided into magnetic
domains separated by the domain walls, and different walls can move
independently producing the average magnetization change (or not). The
Preisach Model and Simulation of Relaxation Kinetics 509

generalized coordinates corresponding to these internal degrees of freedom


are the positions of the individual domain walls. It is clear that there can be, in
general, an infinite number of states (domain walls arrangements)
corresponding to the given fixed value of the macroscopic magnetization.
Moreover, which is less clear and comes from the analysis of the energy of the
whole magnet, the energies of most of these states are very similar to each
other. Thus we can call such states quasi-degenerated; the existence of these
states is usually connected to the microscopic inhomogeneities in the material.
To have an example of a quasi-degenerated states, consider a large, flat
rectangular slab of a soft magnetic material with easy anisotropy axis
perpendicular to the largest surface. Let us assume that such slab contains the
parallel stripe domain structure (there can be other arrangements, which in
fact increases the degeneracy even further), with magnetization in stripes
alternating in the directions parallel and anti-parallel to the easy axis. Suppose
the slab is large enough, so that the boundary conditions do not play an
essential role. If the material is homogeneous, there is a global energy
minimum corresponding to a certain period of stripes. Starting from this
perfect arrangement in a perfect material, consider a configuration with a
single stripe (surrounded by two domain walls) moved as a whole in the
direction perpendicular to its walls. Such a modified arrangement can be held
in place (stabilized) by the inhomogeneities in the material and has exactly the
same magnetization as the original one. The energy is larger than the energy
of a perfect periodic structure, however, if the material is large enough, the
energy surplus is negligible with respect to the total energy of other stripes.
Because any given stripe can be chosen for displacement, in the
inhomogeneous material of the considered structure there is a huge number of
configurations stabilized by inhomogeneities with the energies only slightly
larger than the ground state. These states are quasi-degenerated.
The temperature causes the system to jump between different states.
Because the quasi-degenerated states with lower energy are preferred, the
non-equilibrium system in contact with thermostat decreases its total energy
until it appears in the state of the thermal equilibrium. This process is called
the thermal relaxation. Let us first assume that the temperature is zero and
look at field-induced evolution of the system.

11. 2 The Response Operator

Consider the case of the system described by a certain scalar macroscopic


state variable M ( t), and subjected to the action of a scalar force H ( t), both
functions of time t. As it was already mentioned, in a magnetic material M( t)
can be a projection of the average magnetization vector, and H ( t) a
510 K. L. Metlov

projection of the external magnetic field. Suppose the temperature is zero and
there are no dynamic effects (the experiment is slow enough to let the
dissipation hide them). In this case the response is rate independent, and, in
the most general case, the macroscopic state variable and force can be
related through the response operator f as

M( t) = ,HCt) (11.1)
.
where, is an operator in a sense that it maps a function to a function. In such
a way any rate-independent response can be written. In the case of an a linear
system, the action of the response operator is equivalent to the multiplication
by a scalar. In the case of an equilibrium system, there is a unique
correspondence between the value of the state variable and the force, so that
the response operator is a certain function , ( H), also called the anhysteretic
.
curve. In other cases, is a non-linear operator with memory.
In order to study an operator it is necessary to build its representation,
which describes its action in some way. For example, linear operators in the
functional space are representable by matrices (can be continuous and/or
infinite). The operator , is non-linear; therefore, the theory of linear
operators is not applicable and its representation has to be built in a different
way.
The other problem with , is that this operator has memory, so that the
value of the macroscopic state variable depends not only on the current value
of the force, but also on all its values in the past.
Both non-linearity and memory require a special handling of response
operators.

11. 3 The Preisach Model

The complexity of the response operator lies in the fact that it describes an
essentially infinite-dimensional object. To realize this, suppose, we would like
to store complete response of a system with the rate independent hysteresis. If
we start from a well-defined state (say, saturated state for magnetic systems)
and change the force (magnetic field) monotonously, the state variable of a
system will change along a certain branch. This process maps the values of the
force to the values of the state variable and this mapping can be stored as a
function. But, at any point along this curve this process can be stopped and its
direction reversed, so that the resulting time dependence of the magnetic field
becomes non-monotonous. Such reversals introduce new branches, functions,
which also need to be stored. They are called the first order branches by the
Preisach Model and Simulation of Relaxation Kinetics 511

number of the field reversals required to reach them from the initial state.
There is an infinite number of possible branching points along the original
(zero-order) branch; thus, there is an infinite number of the first order
branches. If we would Iike to store zero and all the first order branches, it is
necessary to be able to store a function of two variables, which is defined on a
two-dimensional space. Along each of the first order branches, the infinite
number of the second order branches can be initiated, which requires adding
another dimension to stored data. Because the branching can be performed at
will any number of times (even infinite), the straightforward exact
representation of a general response operator is an infinite-dimensional object.
In order to build a useful representation with a lower dimensionality, it is
necessary to incorporate additional knowledge of the response operator into
the model. The reduction of dimensionality comes from the fact that additional
general properties of the response operator enforced by the representation
make it possible to store fewer data and to predict the missing using the
properties themselves. From this point starts a broad spectrum of opinions on
how to do it the most general way. All of these approaches are applicable only
to particular classes of physical systems, and, at this time, there is no single,
universal one.
A possible simplification may be derived applying the ideas of equilibrium
thermodynamics, which basically say: "no one cares about particular
microscopic state variables of the ideal gas (coordinates and velocities of
individual molecules), the macroscopic variables (volume, pressure)
describe the state completely." Applied to the above, this results in the
concept of the response with local memory. That is, only the current values of
the macroscopic state variables are important. Once a particular value of such
a variable is reached at a particular value of the force there are only two
possible branches, corresponding to increase and to decrease of the force
from that point. If the assumption of the local memory is accepted, the
evolution of the system can be expressed in terms of the current values of the
macroscopic variables and their infinitesimal increments. Thus, the complete
response (and the response operator) can be represented using the differential
equations in terms of M ( t) and H ( t). A particular form of these equations
(Jiles and Atherton, 1986) and its modifications known as Jiles-Atherton model
is often used (and often very successful) for practical description of magnetic
hysteresis. However, as it was noted above, the main assumption of local
memory is connected with the assumption of the system being at equilibrium,
or with the assumption of statistical independence of fluctuations. Really, in an
equilibrium system the macroscopic variables are fluctuating around their
average values, but the relationship (equations of state) between these
averages at a certain (current) instant of time is established by the
thermodynamical probability distribution function (describing the probability to
find a system in a particular microscopic state). Because the shape of
thermodynamical probability distribution function is fixed and is expressed
512 K. L. Metlov

through the values of macroscopic averages, only their current values matter;
knowledge of all these values completely defines the state of an equilibrium
system. In the absence of equilibrium the probability distribution function is
changing as the system evolves, and is non-stationary even if the external
forces are constant. In this case, the current values of macroscopic variables
do not define the state of the system anymore, because the Iink (the
probability distribution function) is missing (is variable). Thus, the
assumption of the local memory does not play well with the fact that the
hysteresis is mainly the property of non-equilibrium systems with many internal
degrees of freedom (each participating in the definition of the current state of
the system) .
In a system with hysteresis having non-local memory, a large number of
substantially different branches can pass through the point having particular
chosen values of M and H. Particularly, the magnetic hysteresis of domain
walls has non-local memory. The system with non-local memory is impossible
to describe by differential equations in terms of the macroscopic quantities.
The Preisach model (PM) allows us to establish the relationship between
the macroscopic state and the force through non-local integral relation, which
is not reducible to the differential equations. It is based on two assumptions on
the response (Mayergoyz, 1986):
(1) The return point memory, or, the "wiping-out" property. This means
after a complete cycle of the force, the system returns exactly to the
microscopic state before the cycle (the system remembers the return point) ,
or, that the system "forgets" the information about the complete cycles of the
applied force (as if they were completely absent). In systems having the
return point memory the hysteresis loops are always closed. The return point
memory was demonstrated experimentally (e. g., in magnetic hysteresis of
superconductors by Friedman et al., 1994), and also proven rigorously
(Sethna et al., 1993) for systems with order-preserving dynamics. The
system is said to have such type of dynamics if there exists a partial ordering
of microscopic states of the system (a rule allowing us to compare some of the
states) and the ordering between two states is preserved during the evolution
in the same external conditions. That is, having two systems, initially in states
A(O) and B(O), so that A(O)~B(O) according to established ordering, and
applying the same external force H ( t >O)to both these systems, the ordering
stays intact during the evolution, A ( t) ~ B ( t). The partial ordering can be
established rigorously for some model systems, such as the random field Ising
model (RFIM) at zero temperature, or the ensembles of systems in random
potential considered here. Not all systems have the partial ordering (most
notably it is destroyed by thermal fluctuations, as e. g., the partial ordering of
RFIM), there can also be other effects leading to the absence of the return
point memory. The systems with no return point memory exhibit the
"accommodation" effect; the minor hysteresis loops drift gradually when
cycling the magnetic field between two fixed values for extended period of
Preisach Model and Simu lation of Relaxation Kinetics 513

time.
(2) The congruency of minor loops. This means all (starting from
arbitrary initial state of the system) the minor hysteresis loops obtained by
cycling the external force between two arbitrarily chosen fixed values are
geometrically congruent (Fig. 11. 1). This property is less natural for
magnetic hysteresis than the return point memory, and usually (for saturation
to saturation hysteresis) holds only approximately, as there are various tricks
(like e.g. "Moving Preisach Model" approach, Oti et aI., 1991> employed to
transform the hysteresis loops in order to make the congruency to hold more
precisely. It was shown experimentally that some of the hysteresis loops of a
superconductor are to a large extent congruent (Friedman et aI., 1994). It is
also possible to prove congruency rigorously for some systems without
saturation (Bertotti et al., 1996; Bertotti et al., 1999; Metlov, 2000), the
results of the last work will be covered later in the text in greater detail.

~~

~ 0 ---------~+.~~~----------

-1 t:o=:======---_--.l- _
H

Figure 11. 1 Two geometrically congruent hysteresis loops between the same peak values
of the input H marked by two thin vertical lines. The congruency property means that ALL
possible hysteresis loops between ANY peak values of the input are congruent. as shown
in this figure.

While these properties are possible to demonstrate experimentally for


particular minor loops, in order to guarantee the representability they must
hold for all possible loops. Which means that it is impossible to establish them
by a direct experiment. They either have to be proven theoretically to be
present for a particular class of systems or one has to live with partial
experimental checks of these properties. These partial checks can be done
either by measuring a few selected minor loops, or, more comprehensively,
by using necessary conditions of the applicability of the Preisach model
described further in this section.
Both the congruency and the wiping-out properties, if present, allow
514 K. L. Metlov

representing the response operator I in the following way

r- = II
,,>~
dexdf3p(ex,f3) y",~ (11. 2)

where p ( ex , (3) is a function representing the particular response operator,


often called the Preisach distribution function (PDF); Y
".~ is the response
operator belonging to the two-parametric family of primitive response
operators with rectangular hysteresis loops (F ig. 11. 2) Each of the operators
y ".~ remembers one bit of information about its current state, which can be
either 1 ("on") or-1 ("off"). Each of the operators y".~ independently
changes its state from "off" to "on" once the external field reaches the
threshold value ex, and from "on" to "off" when the field is equal to the second
threshold value f3< ex. The value of the response in the current states of the
system described by Eq, ( 11 , 2) is the sum (represented by the integral) of
current states of all y",~ with the weight p (ex, (3) .
M
1

o /3'He a H
-1

Figure II. 2 Illustration of the Preisach model the central plot shows a Preisach
distribution function and the coordinate system, each point on the a, {3 plane corresponds
to a hysteresis operator. The inset shows response of a single operator with switching
fields a, {3 .

It is easy to verify directly, that for any function p ( ex , (3) the response
described by the operator Eq. (11,2) always has the return point memory and
congruency properties. The inverse statement, that any response having both
these properties is uniquely representable by Eq. (11, 2) with a certain
p ( ex , (3), is much less obvious; its proof was first given by Mayergoyz
Preisach Model and Simulation of Relaxation Kinetics 515

(1986). The essence of this proof is that higher order hysteresis branches CD of
the response of a system having the return point memory and congruency
properties can be uniquely expressed in terms of the first order branches.
Particularly, it means that the PDF is completely defined by the first order
hysteresis branches of the system, starting from a well-defined initial state.
Before deriving the formula for the PDF let us consider how Eq. (11. 2)
allows calculating the response M( t) for a particular field history H( t). To do
this, let us first write the response given by the operator Eq. (11 . 2) in the
following form, making use the fact that operators ycx.~ have only two states

M(t) = rH(t) = II
5+ (t)
p(a,{3)dad{3- II
5_ (t)
p(a,{3)dad{3 (11.3)

where $ + and $ - denote the time-dependent sets of operators y cx,~ on the


Preisach plane Tr< ycx,~ En if a> (3) in "on" and "off" states respectively. It
is obvious that the joining of these sets is the Preisach plane in its entirety
$ + U $ - = T7 and that these sets never overlap $ + $ - = ;25. The n
advantage of this notation is that the evolution of the system can be tracked
independently of the PDF by looking on how the sets $ + and $ - change when
the field is applied.
Suppose initially the system is in the state of negative saturation, the field
is H = - 00, so that the sets $ + =;25 and $ - = T7. Now, let us increase the
external field H up to the value Ht , during this process the operators ycx,~
t
with" on" switching fields a < H will switch and move from the set $ - to
$ + , the corresponding configuration of these sets is shown in Fig. 11.3a. The
t
response as a function of the magnetic field H can be written as

H~ 0: a

f f f f
00

MA(Hi) =2 da d{3p(a,{3) - da d{3p(a,{3) (11.4)

where the fact that $ + U $ - = T7 was used and the second integral runs over
the whole Preisach plane T7, the index "A" in M A marks the branch as
ascending. This expression describes zero-order hysteresis branch, because
no reversals took place yet. If we now make a reversal and decrease the field
to the value HI < H t,
the operators with {3 > HI belonging to the $ + set in
Fig, 11, 3a will switch and move back to the $- set; the resulting configuration
is shown in Fig. 11. 3b. The response as a function of both the field when the
t
first reversal had started H = H R and the field H = HI < H can be t
represented as

CD As it was already said, the order of the hysteresis branch M ( t) is the number of
reversals (changes of signs of the time derivative of the force () H/ () t) along it.
516 K. L. Metlov

f3 HI
f3 HI

Cl Cl

S S_

S+ S+

H, H,
(a) (b)

f3 HI

(e) (d)

Figure 11.3 Switching of the operators Y.,P


from the Preisach basis during the evolution
of the system. Differently shaded areas denote the sets of operators in "on" ($ + ) and
"off" ($-) states. The figure (a) corresponds to the increase of the external field up to
the value Hi starting from the state of negative saturation; (b) shows the state of
operators after subsequent decrease of the field to the value HI <Hi; (c) the state after
yet another reversal, increasing the field to H! < Hi; (d) the state appearing after a
series of field oscillations with decaying amplitude.

H~ a

M D (Hi, H,) = M A (Hi> - 2 f f dO' d{3p (0' ,(3) . (11.5)


H, H,
The index "0" in M D notes that the branch is descending, that is corresponds
to decrease of the field. The subsequent increase of the field up to the value
Hi <Ht will lead to the configuration Fig. 11.3c (note that if Hi ~Ht the
configuration will become the same as in Fig. 11. 3a with Ht replaced by Hi .
This illustrates the return point memory property and shows how the
information about the past extrema is erased), the decaying field oscillations
produce Fig. 11. 3d, etc. However, as it is stated by the theorem by Mayergoyz
the first order branches M D ( HR , H < HR) are all we need to restore the
Preisach distribution. It is obvious from Eq. ( 11 .5) that the PDF is given by
Preisach Model and Simulation of Relaxation Kinetics 517

(11 .6)

This representation is two-dimensional thanks to the simplifying return point


memory and congruency properties. The function p ( a , 13) can be also found
from the first order hysteresis branches of the system starting from another
well-defined state. If the system is representable by the PM the PDFs resulting
from the first order branches starting from any initial state must coincide.
Comparing the PDFs from a few selected initial states gives the necessary
conditions for the representability of the system by the PM and might serve as
useful checks before application of the PM to a given set of the experimental
data. An especially simple and useful necessary condition of the PM
applicability results from comparing the PDFs built starting from different
saturated states (e. g., negative and positive ones). Denoting the ascending
first order branches as M A (H R ,H> H R) the expression for the PDF is

(11. 7)

Taking into account that for the usual magnetic material there is no preferred
direction of the magnetization vector except that given by the externally
appl ied magnetic field, allows writing M A ( HR , H) =- MD ( - H R' - H).
Then, comparison of Eq. (11.6) and Eq. (11.7) results in a simple necessary
condition for the representability of the system with a given PDF restored by
the first order hysteresis branches starting from either of the saturated states

p(a,f3) = p(- a, - 13) (11.8)

That is, the PDF of the PM-representable system built from the first order
hysteresis branches starting from a saturated state must be symmetric with
respect to the line a = - 13 on the Preisach plane. Otherwise, the system is
not representable by the PM.
Similar conditions can be built by comparing PDFs for a different set of
initial states, but all of them will be necessary, not sufficient. These
conditions become both necessary and sufficient in the PDFs built starting from
all reachable states of the system (the proof of this statement is obvious).
From the experimental point of view such conditions are not more useful than
complete checks for wiping-out and congruency of minor loops, in the sense
that they cannot be performed completely. However, the weaker, necessary
only variant of these conditions can be applied experimentally and can be a
useful safeguard against attempts to represent evolution of a system with the
PM in the case when it is not representable.
518 K. L. Metlov

11. 4 Ensembles of Systems in Random Potential

The approach we have discussed so far was purely mathematical; no actual


knowledge of the physics of the system (except symmetry with respect to
H- - H, M- - M in one case) was used. The only physical model
employed was a set of non-interacting particles with rectangular hysteresis
loops each, proposed originally by Preisach. Such a physical model is very far
from reality and certainly does not fit for the overwhelming majority of systems
whose response is successfully represented by the PM experimentally. Let us
now concentrate on a question if it is possible to find a more realistic physical
model (besides the aforementioned one) whose response is exactly
representable by the PM.
Let us start from a very general (but one-dimensional) problem. Suppose
when no external force is appl ied, the dependence of the energy of a system
on the generalized coordinate x is given by a function Up (x), which will be
called here the pinning potential. The external force h will be described by the
additional energy term - xh. The total potential for a single such system is
u ( x , h) = Up ( x) - xh. ( 11 . 9)

The equilibrium values of the coordinate x at a given value of force satisfy

h = h p (x), h p ( x) = dup(x)
dx . (11.10)

For non-linear systems (in non-parabolic pinning potentials Up (x there can


be many solutions of Eq. (11 . 10). The quasistatic evolution of the system can
be obtained as the limit dh ( t) / d t-O of the following dynamic equation

(11.11)

This equation is non-linear and does not assume that the system is close to
equilibrium. Because we are now interested in quasistatic evolution, the
higher order time derivatives were omitted from Eq. ( 11. 11). The evolution
proceeds by reversible motion in between the local extrema of h p (x),
followed by jumps (when an extrema of h p (x) are reached) to new regions of
the reversible motion (Fig. 11.4).
Now, if we try to represent the response x as a function of h by the PM,
we will soon discover that it is possible only for very few and unrealistic
potentials Up ( x ). The wiping-out property is not a problem, it holds for
dynam ics (Eq. ( 11 . 11 in any continuous potential Up ( x) (Bertotti, 1998).
The congruency, however, requires, besides other things, that all the parts of
the potential corresponding to the reversible motion are parallel to each other
Preisach Model and Simulation of Relaxation Kinetics 519

o
hi --- _-H-+------ .JI

-0.05'-----L~__::'_--'----------'-------'--
XJXL 350 400 450
x

Figure 11. 4 Evolution in a single realization of a random potential (a single system in the
ensemble) generated by the Campbell process (Eqs. ( 11 . 20), (11. 21>, (11. 22, with
the illustrative values of the parameters. The thick solid line shows a realization of
hysteresis branch corresponding to the field history h ( t) : = 00 - - hi -- h 2 - - h 3 The
coordinates x J and XL mark the beginning and the end of one of the jumps.

in the regions of fields h p where they overlap (because the congruency, in


particular, means that the hysteresis branches are parallel). This condition for
the potential of a single system is very unrealistic; it is satisfied for piecewise
quadratic potentials Up (x), but not for many others.
The situation becomes much better if we consider the ensemble of
systems in a potential Eq, ( 11 . 9) with different Up (x) in each system being a
realization of a random process. It turns out that the ensemble-averaged
response x can be exactly representable by the PM. Bertotti and coworkers
performed the first step in this direction by proving that the ensemble of
systems, each in a random potential being the Wiener-Levy random process,
with independent increments is exactly representable by the PM (Bertotti et
aI., 1996). The resulting PDF belongs to the class of non-saturable ones; it
depends only on difference of its arguments p ( a , f3) = p ( a - f3) and the
response resulting from such representation grows unbounded when the
external force is increased infinitely (but it shows a non-local memory).
The most common physical situation suitable for the modeling as the
ensemble of systems in a random potential is an array of magnetic domain
walls in soft magnetic material with inhomogeneities. At sufficiently small
magnetic fields (tar from the fields of domain wall nucleation/annihilation), the
magnetization process in such material proceeds through the domain wall
motion. It is possible to associate a single system in the ensemble with a
520 K. L. Mellov

single domain wall in the array, and the potential felt by that single wall is
Up (x). Different walls in the array feel the different interaction with the
material, but assuming, naturally, that the material is uniform (but still
inhomogeneous), it is clear that statistically all these different Up ( x) are
similar, or. are realizations of a single random process. Different wall
structures. their curvature, and inhomogeneities types influence the details of
the random process for Up (x). The magnetization of such an array is
proportional to the average displacement of all domain walls. or to the
ensemble average x.This concludes the analogy and allows us to think about
the problem discussed below in terms of the simple domain wall motion.

11. 5 Representability of the Ensemble Evolution by


the PM

Let us now address the question: What properties must the random process for
Up (x) obey in order for the average response x to be representable by the
Preisach model. This question was considered in two recent works. First
(Bertotti et ai., 1999) it was proven that any homogenous Markovean process
generates the evolution representable by the PM. Shortly after that (Metlov,
2000) the representability was proven for a wider class of random processes
including the non-Markovean ones. In this chapter the non-Markovean
potentials will be considered. The Markovean processes are not very realistic
in the case of the domain wall motion because if they have the derivatives.
these derivatives are, at most, discontinuous. On the other hand, the
potential a single domain wall feels in an inhomogeneous material (does not
matter how jagged the inhomogeneities are) is always smooth (Eq. (11.22)
and commentary afterwards) because of the spatial structure (and non-zero
thickness) of the domain wall. which smoothes the inhomogeneities.
Let us again consider a system in a particular realization of the random
process for Up (x) and calculate a branch starting from some metastable state
x = x 1 at a given value of the external force h = hI when the force changes
monotonically to the value h = h 2 . This branch can be found taking the
quasistatic limit of the solution of Eq. ( 11 . 11) with the initial conditions listed
above, or, equivalently by following the rules (see Fig. 11.4 for illustration) .
( 1) The system follows the pinning field (x is the solution of h p (x) = h)
until (condition 1) it reaches a jump point x J (where h; (x J) = 0) .
(2) The system jumps (x changes infinitely fast up to x = XL) until
(condition 2) the next closest metastable state (the root of h p (x L #- x J) =
hp(xJ'
These steps are repeated until the desirable value of the final value of the
force, h 2 , is reached. They define the unique hysteresis branch starting from
Preisach Model and Simulation of Relaxation Kinetics 521

a given metastable state x 1 of the system in the ensemble. Let us denote this
>
branch h ~ (x, Xl)' where sign depends on whether h z hI or not. It is more
convenient to consider the force h as the function of the coordinate x because
for all points inside the branch this function is single-valued and has a finite
derivative (if h; (x) is bounded). It is even more convenient to represent
x

h~(X,Xl) = hI
x,
f
+ h~'(x)dx (11.12)

and to reformulate the above rules for h ~. (x , x, )

h~' (x) = {h; 0,


(x) , until condition 1 is satisfied
until condition 2 is satisfied
(11.13)

where arrows denote cycling of conditions. Please note that the branch is
calculated starting from a turning point of the field. Only then we may assume
that the first infinitesimal change of the coordinate will be reversible and start
looking for condition 1 to be satisfied first, as implied by the left arrow
pointing upwards.
The representability of the ensemble-averaged hysteresis generated by a
particular random process for Up (x) can be established by looking for its
wiping-out and congruency properties. It can be proven (Bertotti, 1998a),
using the general recipe of the proof by (Sethna et aI., 1993), that the wiping-
out (the return point memory) property holds for all systems in the ensemble
evolving according to Eq. (11 . 11) individually (and, thus, for the ensemble as
a whole).
The congruency of the minor loops with peak fields hI = h p (x 1) and h z
impl ies that the field increment h ~ (x , XI) - hI averaged over the ensemble of
systems in different real izations of Up ( x) along the branch of the loop, is a
function of the position increment x - x 1 only
x

h~(X,Xl) - h, f
= h~'(x,xl)dx =
x,
f(x - Xl) (11. 14)

and is independent on the state at one of the turning points of the loop.
There is a well-known property (e. g., Korn and Korn, 1968) of averages
of any stationary random process p ( t), namely, the following relation holds
for the average over an ensemble

q(pCtI),pCtZ)"",pCt n = q(p(O),pCt z - t 1 ),,pCt n - t1


(11.15)

for any function q. It means the average depends on n - 1 difference of


arguments only, leaving one of them as an independent variable. The
similarity of the Eqs. (11. 14) and (11. 15) proves the congruency if:
( 1) h; (x) is a stationary random process,
522 K. L. Metlov

r
(2) h (x, X 1 ) depends only on h; ( x) evaluated at a number of
points x.
The latter can be shown constructively. Indeed, the body of Eq. ( 11 . 13)
and the "condition 1" already depends on h; (x) only. The "condition 2" for
the coordinate of the end of jump x L can be rewritten as
XL =;t=.X J

f h;
X
(x)dx =0 (11.16)
J

which now depends on the spatial derivative of pinning field h; (x) only,
evaluated at a number of coordinates. It means the whole branch can be
expressed (in a very complex way, but the expl icit form of this dependence
does not matter here) as a function of the values of h; (x) only, evaluated at
a number of coordinates. Thus, if h;
(x) is stationary, one of coordinates
becomes free after averaging over the ensemble, and the last equal ity in
Eq. (11 . 14) holds. This proves the statement: if the first derivative of the
pinning field h; (x) is a stationary random process, all average hysteresis
loops between the same peak values of the external field are geometrically
congruent.
Thus, the congruency and the wiping-out property are satisfied for the
average evolution of the ensemble of systems in a random potential Up (x)
whose second derivative (or the first derivative of the corresponding pinning
field Eq. (11. 10 is a stationary random process. Consequently, the
stationarity of u~' (x) is sufficient to guarantee the representability of the
average evolution of the ensemble by the Preisach model.
Turning back to our domain wall motion analogy, the requirement of the
stationarity means that the statistical properties of the inhomogeneities are the
same, independently of how far each individual domain wall in the array have
moved. This means, by moving all domain walls by a certain distance
synchronously (say, by "demagnetizing" in a bias field) the response when
varying the field around the bias is independent on the bias except for a certain
overall constant displacement. This necessarily implies that the Preisach
distribution function has a non-saturable form p (ex, f3) = p ( ex - f3), as this is
the only form of PDF producing the response invariant (forgetting about the
overall shift) with respect to the force bias.
Another interesting fact is that adding the parabolic term to each
realization of the potential up(x)-up(x)+gx 2 /2 does not destroy the
representability of the ensemble response because its second derivative is a
constant in up (x). That is, if the evolution in the original random potential
up (x) is representable by the PM, the evolution of a stab iIized ensemble in
the potential up (x) + gx 2 /2 is also representable. Higher order terms in the
potential may (or may not, because the above discussion gives a sufficient
condition of representability only) make the evolution not representable by the
PM. The higher order terms introduce the saturation in the response and the
Preisach Model and Simulation of Relaxation Kinetics 523

PDF describing it will depend on two arguments expl icitly.


A single first order branch is sufficient to restore such a non-saturable
PDF. This allows us to conclude from the previous discussion: a system far
from saturation under the natural assumptions of statistically uniform
inhomogeneities must be exactly representable by the Preisach model and fully
characterized by a single first order hysteresis branch. While the arguments of
this section apply to one-dimensional evolution only, this statement is probably
true for multi-dimensional systems as well (but this requires a separate proof,
currently not done), as it would explain a lot of experimental successes in
representing the evolution of such systems by the Preisach model described in
literature 0 .

11. 6 Connection of the "Classical" Irreversibility


Parameters with the PM

Conventionally, in studies of magnetic hysteresis of domain walls, the domain


wall coercive field Hew parameter is used to characterize the irreversibility.
The Preisach model provides a richer picture, but also allows expressing the
conventional domain wall coercive field parameter through the non-saturable
PDF. The formula will be useful for our further discussion.
The Hew parameter is usually defined (Pu st et ai., 1996) as the value of
the magnetic field H corresponding to the point of linear extrapolation of the
magnetization curve M ( H) at large values of H starting from demagnetized
state crosses the line M = 0 (Fig. 11. 5). The initial magnetization curve can
be expressed using the Preisach model with non-saturable distribution function
p ( cx /2 - 13 /2) = p ( He) as
H 0: H HI

M(H) = 2fdCX f d13P(CX; 13) = fdHlf dHep(H e) (11.17)


o -0: 0 0

where HI = (CX + 13) /2 and He = (CX - 13) /2 are just convenient variables of
integration over the Preisach plane in the case of non-saturable PDF. Provided
the integral

(11.18)

has a finite value, which is true if p(H e ) decays sufficiently quickly as He-

o One can look up, e. g., the index of IEEE Transactions on Magnetics for the
keywords "Preisach model" to find many examples.
524 K. L. Metlov

~ 3

o 6 8 10
H

Figure 11. 5 The definition of the domain wall coercive field. The dashed straight line is
tangential to the initial magnetization curve (solid line) at infinity (H-~ 00). The fact that
these lines do not look tangential in the range of fields shown illustrates the point that such
a definition of the Hew when applied experimentally always underestimates its value. The
parameters are defined in the text, the Preisach distribution function p ( H) e = H; / ( , +
H~) was used in this example.

00, the domain wall coercive field can be expressed as


~

~ H, f HcpCHe)dH e
Hew = _1 fdH1[Xs- fdHePCHe)J= -,- : "o~ C11. 19)
fpCHe)dH e
_

Xs 0 0

o
where the last equality is due to integration by parts. Thus, the domain wall
coercive field Hew is just the average He with p CHe) understood as a
probability distribution function. The parameter X s is proportional to the initial
susceptibility of the system measured after it is equilibrated by oscillations of
the external field with the amplitude slowly decaying to zero, and it is also the
slope of the anhysteretic curve, which is a straight line for systems described
by the non-saturable PDFs.
Equation C11.19) involves integrals with infinite limits, which is the tricky
point worth discussion. From the mathematical point of view these integrals
must converge Cwhich means the pC He) must decay faster than 1/ H~ at large
values of He -- 00 ). In real ity, however, the non-saturable PDF for a given
system is not defined for large values He because of the non-linearities
connected with the domain structure hysteresis, which lead to saturation.
Depending on the Preisach distribution function the finite range of non-saturable
PDF Cor initial magnetization curve) may be Cand may not be) enough
depending on how fast the integrals in Eq. C11. 19) converge. The error
connected with the finiteness of the measurement field range is visible in
Preisach Model and Simulation of Relaxation Kinetics 525

Fig. 11.7. Really, given the range of the initial magnetization curve shown in
the figure, the experimentalist would certainly draw a different tangential
straight line than the one shown (which is exactly tangential at H-oo), and,
consequently, obtain a value of Hew different from the exact one used to draw
the figure. This suggests the necessity of introducing the finite cut-off field H cut
into integrals in Eq. (11.19) by replacing the infinity signs there with H eut This
would also allow extending Eq. (11.19) to the PDFs for which the integrals of
Eq. ( 11. 19) do not converge. The non-convergence can be caused by the
behavior of PDF at large values of He' which can never be physical because of
saturation effects. For all PDFs, except the one for Campbell, random
processes built and discussed below the integrals of Eq. (11.19) do converge.
For the Campbell process, the qualitative dependence of Hew on 9 discussed
here does not depend on the value of cut-off field once it is reasonably large.

11. 7 Representations of Some Ensembles by the PM

To find the PM-representation (the PDF) of evolution in a particular random


process with a stationary second derivative (so that the PDF is non-
saturable), it is necessary to find a single first order average hysteresis
branch. The problem of finding such a branch is equivalent to solving the so-
called "exit problem" of a theory of random processes, and can be solved
explicitly (and in a general case) only for Markovean random processes. The
PDF for such random potentials was given in Bertotti et al. (1999). For non-
Markovean random processes there is no general solution of the "exit
problem" and, consequently, the representations need to be built for each
class of potentials separately. Several classes of potentials will be considered
in this section.

11.7. 1 Campbell Random Potential and Stabilization of


Domain Walls
Let us first study the influence of the stabilization of individual systems on the
response. Consider the ensemble with the following pinning potential
(11.20)

and take the Campbell random process to generate Up (X), so that

(11. 21)

where a i are the Gaussian distributed random numbers, and X i are positions
526 K. L. Metlov

of inhomogeneities. The X;-s are generated by a Poisson random process, so


that random quantity X; - X ;-1 is distributed according to the Poisson
distribution exp X; - X ;-1 ) / 0 ) / 0, where 0 is an average period of
inhomogeneities. The function E (X) describes the interaction of a domain wall
with a single inhomogeneity; it depends on the particular problem that is
solved.
For example, let us take the 180' straight Bloch domain wall having the
profile If 0 (X) = 2arctan ( e X/6 w ), where If, is the deviation of the
magnetization vector inside the wall from its direction in the neighboring
domains (which is If 0 (x- - 00 ) = 0 and If 0 (x- 00 ) = n) ,X is the coordinate
perpendicular to the wall, ow=n JA/K is the wall width, A and K are the
exchange and the uniaxial anisotropy constants of the material, respectively
(Fig. 11. 6). The interaction of this wall with the planar rectangular
inhomogeneity of the anisotropy constant of the width 0< S; Ow and the small
amplitude K; = l:l.K j / K, as shown in Fig. 11. 5 results in the following
interaction energy density per unit of wall area
E't'(X - X;) =

O'w [ 1+ n; j (tanh (X - Xs~ + S ;/2 - tanh (X - Xs~ - Sj /2) ]


(11.22)

1.0 /,
I
I Ow II

~I
I
I
0.8 I
I
I
I
1'-
I
Sj
.. ,
I
J
I
I -------
I
~0.6

-
J
'X
~
;:; 0.4
Ix;
0.2
/1/
I
1/
o o
X
Figure 11. 6 The 180' Bloch domain wall interacting with the rectangular inhomogeneity of
the anisotropy constant. The variables are defined in the tex.

This expression can be obtained in the following way. First, because the amplitude
on anisotropy constant inhomogeneities is assumed to be small, we limit consideration to the
linear approximation over this amplitude K i (but not over the domain wall displacement). In
the linear approximation the distortions of the wall profile due to inhomogeneities can be
neglected. Thus, Eq. (11. 22) is just an integrated energy density of inhomogeneous
anisotropy K ( x) sin 2 ( If' 0 ( x calculated starting with the undistorted Bloch domain wall
profile.
Preisach Model and Simulation of Relaxation Kinetics 527

where C7w = 4 ~ is the domain wall energy density in the absence of


inhomogeneities, X is the position of the domain wall and X i is the position of
the inhomogeneity center, An interesting feature of this potential is that while
the spatial profile of the anisotropy constant has the discontinuous derivative,
the interaction potential is smoothed by the domain wall and has continuous
derivatives up to any order, If, for simplicity, we assume that the width of
anisotropy constant inhomogeneities S i is small, and keep only terms of the
first order in S i we get

E w1 ( X X) - {1 + 1T K is i 1 } (11.23)
- i - C7 40 w cosh 2 [1T(X - X)/o] ,

From this expression it is clear that once the increment of the anisotropy
constant in the inhomogeneity is positive ~K i , a i > 0 the interaction of the
domain wall and a single inhomogeneity is repulsive (wall tries to avoid being
positioned at the inhomogeneity), and in the absence of other inhomogeneities
the wall will move away to the infinity. For the purpose of calculating the
evolution of the system we omit the constant in E (X - Xi), which is the first
term in Eq. ( 11 . 23), and also put a i = K i ' which can be done without loss of
generality, This gives

E (X - Xi) - 1T K is jC7 1 (11.24)


- 40 w cosh2 [1T(X - Xj)/o]'
It is worth noting that even though the interaction potential produced by each
inhomogeneity is repulsive (i. e., corresponds to the energy maximum), the
total potential of a large number of randomly positioned inhomogeneities has
many local energy minima, as shown in Fig, 11. 4. In this case the potential
for a single wall interacting with many thin (S i
ow) inhomogeneities
(Eq, ( 11 , 24)) at once is given by a real ization of the random process
(Eq, ( 11 , 21 ) ). Different real izations of Eq, ( 11 . 21) (and consequently
Eq. ( 11 ' 20)) correspond to different domain walls in the domain structure,
which is represented by the whole ensemble. The coordinates of each wall in
the structure are independent; each of them moves in its own real ization of the
random potential, but all the walls are subjected to the same external field H,
The role of the quadratic in X term in Eq. ( 11, 20) is to provide the
stabilization of the walls in the structure at their equilibrium positions, This
stabilization is due to the domain structure and is not related to the
inhomogeneities. If we forget about them and put a j = 0, the walls wi II be

This is clear because the anisotropy energy density is K (x) sin2 ( '1'0 (x, the
second multiplier in this expression is always positive, so local increase in the anisotropy
constant means increasing the total energy. Consequently, the system tries to accommodate
to such a configuration when the sin 2 ( '1'0 (x multiplier in the location where the anisotropy
is increased is zero, this corresponds to the wall moved infinitely far from inhomogeneity.
528 K. L. Metlov

situated at their equilibrium positions within the domain structure, defined by


the repulsion of the walls by the magnetostatic interaction within the domain
structure. Because we consider only straight walls, the corresponding
structure consists of parallel stripes. Such a structure is very common in thin
films of magnetic garnets, and in many other magnetic materials. There is a
definite period of stripes, which is defined by the magnetostatic interaction
(Koey and Enz, 1960). Thus, the walls are equally spaced (let us say their
coordinates are X~ = nOs' Os is the period of stripes, n is an integer). If we
displace any of the walls (say, i-th) from its equilibrium position X~ to the
new position X; the force will arise, which will try to pull the wall back. For
small displacements I X~ - X; I Os of domain walls from their equilibrium
positions, this force can be expanded into the Taylor series, which leads to
the additional quadratic term in the potential G ( X; - Xn 2, as in Eq. ( 11 . 20) .
The linear term in the Taylor expansion is absent because the stripe structure
with non-displaced walls X; = X~ corresponds to the global (when no
inhomogeneities are present, a; =0) energy minimum. The values of G were
calculated theoretically for several simple domain structures in (Tomas et al.,
1995) and are directly related to the magnetic susceptibility of the equilibrium
structure (Tomas, 1990). This picture could be applied to the ensemble
Eq. ( 11 . 20) if we measure the displacement of the domain wall in the
ensemble Eq. (11 .20) starting from its equilibrium position in the structure, by
letting X = X; - X~. This is convenient, since the magnetization of the
equilibrium stripe domain structure is zero, and, thus, the average
displacement of the domain walls from their equilibrium positions X is
proportional to the macroscopic magnetization of the whole stripe domain
structure.
Let us now find the dependence of this displacement on the external
magnetic field X ( H ). The Campbell random process (Eq. (11. 21 is
stationary (as well as its derivatives). Therefore, according to the previous
section, the non-saturable Preisach model (Fig. 11. 7) is appl icable to
describe X (H). The calculation of the distribution function itself is still quite a
compl icated problem. It was done in Metlov et al. (1999) for the case of
dense defects Oow, which allows approximating the random process by a
Gaussian one (thanks to the Central Limiting theorem of statistics) to solve the
exit problem. The resulting normalized distribution p(a/2- f3/2) = p(H c )
can be expressed parametrically as

Hc(w,g) = H s [)log(l + w) + wg], }


(11.25)
P ( w,g ) -_ ---,2::=--,g~(,--:::1_+~w---,-)-=-[
~
-'.-1:-+--=27.-lo=:g=;(~1=+=w~)]~
[1 + 2g(1 + w) )log(l + W)]3

with the parameter wE[O,oo),g=rr )5/7 Gow/H s ' where Hs=rrH A


Preisach Model and Simulation of Relaxation Kinetics 529

1.5

.S:; 0, g=10o
.!OJ
C, g=0.275 45 ...
'5 I, g=10-1
'" 1.0
'6 2, g=10-2
..c 3, g=10-3
g
'" 4, g=1O-4
~ 5, g=lO-s
] 0.5
"a

o
Z

o 2 3 4 5
Normalized coercive field, hc=HJHs

Figure 11. 7 Preisach distribution functions for a system of independent Bloch domain
walls interacting with random rectangular inhomogeneities of the anisotropy constant at
different values of the stabi Iizing gradient g. The curve marked by "0" corresponds to
crossover between the PDFs with and without peak .

.J (1T S7KD I (Dow) is the field of start@ of "free" domain wall motion in the
absence of the gradient (G = 0), and H A = KIM s is the anisotropy field.
These distributions are plotted in Fig. 11. 6. It is useful to discuss their
dependence on the gradient g. As it was said above, the gradient term in
Eq. (11.20) describes the stabilization of the domain walls on their equilibrium
positions within the domain structure. For large values 9 the walls are very
well stabilized and difficult to move. Consequently, the field-driven evolution
of the system is dictated by the competition of the stabilizing force and the
push created by the external field. The inhomogeneities do not playa major
role, thus, the whole evolution is very close to being reversible. The
corresponding Preisach distributions are concentrated in the region of small He
and are monotonously decaying (F ig. 11. 6). The small values of the gradient
correspond to nearly free (unstabilized) walls. In this case, the gradient
becomes unimportant, and the evolution starts to be mainly defined by the
competition of the random pinning force and the force produced by the external
field, thus, becoming more irreversible. The Preisach distribution shifts into

@ The field of start of "free" domain wall motion is the value of the external field at
which the inhomogeneities of the material cannot hold the domain wall in place anymore. For
random inhomogeneities (which may have arbitrarily large amplitude and exert arbitrarily
large pinning force upon the domain wall, although with small probability) the wall can never
be completely free, nevertheless, its displacement, while not infinite at H>H s , grows very
rapidly (in the considered case exponentially) with the external field. The parameter with the
dimension of field, setting the scale for this growth plays the role of the field of start in this
case.
530 K. L. Metlov

the region of large He and starts having a maximum (corresponding to the most
numerous jumps in the system). For the considered simple model there is a
well-defined crossover between these two regimes, and the corresponding
values of the gradient and the Preisach distribution shape are marked by the
symbol "C" in Fig. 11. 6.
A more conventional point of view on the dependence of the evolution on
the value of the stabilizing gradient is to look at the domain wall coercive field
parameter Hew, defined by Eq. ( 11. 19) as just at average value of He with
p(H e ) understood as probability distribution function. It is easy to see that in
the limit g- 00 the PDF Eq. (11.25) transforms into Dirak' s delta function
exactly. It is completely concentrated in one point (with He =0) and the area
under it is one. As a result, the Hew decreases (up to zero, in the limit g-
00) with increasing the domain wall stabilization g. This fact is also confirmed

both theoretically and experimentally without using the Preisach model


(Tomas, et aI., 1995).
In practice, the value of stabilizing force acting on the domain walls can
be tailored by creating different domain structures (out of a set of possible
metastable ones) in the material, changing the equilibrium parameters of the
structure (say, by tailoring anisotropy of the material via the temperature and/
or mechanical stress), or by applying a bias magnetic field (so that the
evolution around that bias is modified). These cases and theoretical
calculations of the gradient were considered in Ref. (Tomas, 1990; Tomas et
ai, 1995). In these works also the experimental dependence of the domain
wall coercive field on the gradient was measured.

11. 7. 2 Periodic Potentials with Random Phase


A simple type of non-Markovean potentials allowing us to obtain the Preisach
representation of evolution of corresponding systems in closed form is a family
of periodic potentials with a random phase. Their specialty is the "coherence"
of unreachable states, which will be discussed in the next section. The
corresponding random potential is
u(x) = OH eO W (X /0 + r) + GX 2 - HX (11.26)
where w ( y) is a periodic function with period 1, so that, 0 is the
characteristic spatial scale of the fluctuations, H eO measures their strength, G
is the curvature of the large-scale parabolic energy well, r E [0, 1] is a
uniformly distributed random number. In this section the dimensionless
variables x = X /0, 9 = GO / H cO' h = H / H eO (distinguished by the case of
corresponding letters) are used. The second derivative of the potential
(Eq. (26)) is a stationary random process, thus the evolution of the
corresponding ensemble is exactly representable by the PM. Note however
that the theorem given at the beginning of this chapter covers a much wider
class of potentials than Eq. ( 11 . 26) .
Preisach Model and Simulation of Relaxation Kinetics 531

To track the evolution of individual systems of the ensemble, let us use


the concept of the pinning field h p (x) = [U' (X) + HJj H cO = w' (x + r) + 2gx.
Then, the quasistatic evolution of a single system of the ensemble X ( t) can
be calculated by taking the quasistatic limit h' ( t) -+ 0 of the solution of the
equation of motion Eq. (11 . 11)

(11. 27)

with the initial condition specifying the coordinate of the system at the
beginning of the branch. Consequently, the first order return branch of a single
system of the ensemble is a quasistatic limit of the solution of the following pair
of equations (corresponding to increasing and decreasing branches
respectively)

A d;t~ = ha - h~[xa(t)J, x~(to) =- 00 1


d t~{3
/\., dx = h a{3 - h Pr [ Xa{3 ( t) J , t1 < t< t 2 , X~{3(tl) = x~(tl)f
(11.28)
where h a( to) = - 00, h a ( t 1) = 0' = h a{3 ( t 1) , h a{3 ( t 2) = {3. After taking the
quasistatic limit (h; ( t) -+ + 0, Ii ai t) -+ - 0) the solution for the first order
branch x~{3 does not have an explicit dependence on time t and the damping
factor A. These branches are different for different systems of the ensemble as
facilitated by the upper index r. It is easy to check that both these equations
and their boundary conditions when written for the unknown function x + r
instead of x are equivalent to the equations of motion of a single system having
r = 0 with the reversal fields 0' 1 = 0' + 2 gr, {31 = (3 + 2 gr. This allows us to
write the dependence of x~{3 on r explicitly

X~{3 = Xo(O' + 2gr,{3 + 2gr) - r (11 .29)

where Xo (0', (3) = X~{3 is the first order return branch of the realization with r =
O. The function X~{3 is periodic in r with the period one (x~tO = x~tO) due to
the periodicity of the potential (however, generally speaking, x ~i30:oF x ~;o ).
From this periodicity in r it follows that XO (0' + 2 gn ,{3 + 2 gn) = x 9 (0' , (3) + n
for any integer n.
The average first order return branch x cr{3 can be found by integrating
Eq. (11.29) over r E [0,1], which leads to
Xa{3 = l:i.x (0' - (3) + Xa ,
2g+c

l:i.x(l:i.h) = 2~ f [xO(~,~-l:i.h)
c
- xO(~,~)Jd~
( 11. 30)
2g+c

Xa = 2~ - ~; = 20'9 - 2~ [ c + 9 - f
c
XO (~, ~)d~ ]
532 K. L. Metlov

where hew is the coercive field of the ensemble (as defined in the previous
section), c is an arbitrary constant, and its value can be chosen for
convenience (the result does not depend on it because when integrating a
periodic function over the period the integration region can be freely shifted) .
This expression is valid for any random potential of type (Eq. ( 11.26. The
corresponding PDF can be found using the Eq. ( 11 .7).
The usefulness of the PM manifests itself in the fact that once the PDF is
found it is very easy then to calculate the higher order hysteresis branches and
the response to any quasistatic sequence of external fields without solving the
equations of motion (see Eq. ( 11 . 27) ). The PDF resulting from Eq. ( 11 . 30) is
called non-saturable because the increment x,,~ - x" b.x
= (ex - f3) (as well as
the PDF) depends only on distance from the reversal point b.h = ex - f3.
The other specialty of the periodic potentials with a random phase is that
b.
the first order branch x,,~ at a certain distance h E from the reversal point
exactly coincides with the decreasing envelope branch - x-",
as opposed to
approaching it asymptotically. Since both of the envelope (Eq. (11. 30
branches are straight lines, this means p ( ex - f3 > b.h E) = O. The distance
b.h E is given by
b.h E = max hp(x) - min hp(x) - 2g. (11.3D
O<x<1 O<x<1

It is easy to check that when the evolution is completely reversible, or, there
are no local extrema of h p (x) on the period the minimum and maximum
values are achieved on the boundaries x = 0, 1 and their difference is exactly
2g, thus b.h E = 0 in this case, and, similarly to the limit g- 00 of the PDF
corresponding to evolution in Campbell random process discussed before, the
PDF for the evolution in periodic potentials with the random phase is wholly
concentrated on the line ex = f3 (it is usually said that the values of PDF lying
on this line represent the "reversible" component of the response).
The explicit formula for the PDF itself is easy to write when the potential
has only a single kind of jump per period, or, equivalently, all jumps have
exactly the same length. Then, the evolution of the ensemble is completely
reversible after the reversal point until the opposite envelope is met. Because
the reversible part of the non-saturable PDF depends on ex (or ex + f3) and the
whole non-saturable PDF depends only on ex - f3, the reversible part of the
non-saturable PDF must be constant, and, thus, purely reversible branches
must be linear. This linearity allows to represent b.x
(b.h) = XA b.h. The
constant XA can be found from the condition that the return branch meets the
envelope at b.h = b.h E , which gives
(11.32)

The mechanism of disappearance of local minima in the potential (Eq. ( 11 . 26 is


similar to disappearance of local minima of the function sin (x) + k x , when the value of
k>l
Preisach Model and Simulation of Relaxation Kinetics 533

This formula was tested numerically by evaluating ilxCilh) from Eq. Cl1.30)
for sample potentials

Cl1.33)

and
- - 1 - - 1 - -
wzCy) =- 2yCl + y) - 4C1 - 4y)ZeCy - 1/4) + 4C1 + y)ZeCy + 1/4),
Y= y - intCy) Cll.34)
where eCx) is the unit step function ceCx>O)=l,eCx<O)=O) and intCy)
gives the integer closest to the real number y. Despite looking rather
complicated the potential Wz Cy) corresponds to nothing more than a saw-tooth
pinning field. The average first order branches are shown in Fig. 11.8 and can
be expressed through the parameters hew and hE as
- cx - h
xCcx,(3) = 29 eW - Ccx - (3)X R e[ilh E - Ccx - (3)J

- (_cx - ~9- hew)e[Ccx - (3) - ilhEJ Cl1.35)

Figure 11. 8 Hysteresis branches of a system in a periodic process with a random phase
having a single type of jump.

which results in the Preisach distribution

pCcx,(3) = 2~[ (1 - ~h;nc5CCX - (3) + ~h;:c5Ccx - (3 - E


ilh ) ]
C11.36)
534 K. L. Metlov

where 5(x) = f{(x) is right-sided(1) Dirak's delta function. This PDF allows
calculating the quasistatic response of ensembles with a single type of jump to
any quasistatically applied sequence of external fields. The dependencies of
h cw and lJ.h E on 9 for two particular choices (Eqs. ( 11 . 33) and (11. 34 for
the periodic function w ( x) can be calculated analytically based on
Eq. (11.37) and are shown in Fig. 11.9, but the details of this calculation will
be published elsewhere. For verification of the PDF (Eq. (11.36 one could
substitute it into the Eq. (11. 19) and see that the resulting value is indeed
hew.

N O.6

::a.'"
~ 0.4

0.2

o 0.2 0.4 0.6 0.8 1.0

Figure 11.9 Coercive field hew (upper curves) and the reversibility interval fJ.h E (lower
curves) for quasistatic evolution in potentials having sinusoidal (solid) and saw-tooth
(dashed) pinning fields, QMAX is the value of the effective gradient (1f and 2 for potentials
w, and W2 respectively) at whichevolution of the system becomes completely reversible.

The PDF is applicable to any periodic potential with a random phase, but
depends in a non-trivial way on the particular shape of the periodic function
w (x) through the values of hew, lJ.h E and of the gradient g. The general
dependence of this Preisach distribution on 9 is similar to what was discussed
in the previous section (that is, there are progressively more reversible
components as 9 increases). However, in contrast to the distribution
(Eq. (11.25, for periodic potentials with the single jump there exists a

(1) The right - sided Dirak' s delta function is defined by J5


o
(x) dx = 1 for a > 0,

whereas for normal (centered) delta function such integral is equal to 1/2. It is more
convenient to use right-handed delta functions in the expressions for PDF because its
integration is often going from the line cr = {3 (which corresponds to zero of the argument of
the delta function describing the reversible component of PDF) and the right-sided delta
function makes it unnecessary to keep track of an extra numerical coefficient.
Preisach Model and Simulation of Relaxation Kinetics 535

certain value of the gradient gMAX' such that when g> gMAX the evolution of the
system becomes completely reversible and the peaks on the PDF merge. gMAX = TT ,
2 for potentials (Eqs. ( 11 . 33) and (11. 34 respectively.
It can be expected that if, instead of a single type of jump in the system,
there is a variety of jumps of approximately the same length with small
variance, the peaks on PDF will be broadened (instead of being infinitely
sharp delta functions), as is approximately demonstrated by Eq. (11. 25) ,
where the characteristic length scale of jumps is the domain wall width. Thus,
Eq. (11.36) may be a starting point for interpretation of PDFs with peaks.
An interesting fact is that not all the details of the periodic potential
contribute to the shape of the PDF. Indeed, the potential enters the formula for
the first order branch (Eq. ( 11 . 25 through the function XO (ex, (3). which is
the solution of the dynamic Eq. (11.28) where r = O. or for a single system of
the ensemble. There are certain possibilities to vary the potential of that
single system, which leave XO ( ex , (3) unchanged. This point is relevant for the
next section.

11. 8 Uncertainty in Prediction of Relaxation Kinetics


Based on the PM

As it was said before there are parts of the potential that are unreachable for
quasistatic evolution. The simplest illustration for this fact can be made by
looking at evolution in the following potential
u(x) =-2x 2 +x 4 -hx (11.37)
which has for the field h = 0 two wells of depth 1 situated at x m = 1. The
metastable states of this potential at a given value of the external field hare
given by the equation h p (x) = 4x (x 2 - 1) = h. There is one metastable state
for I hi> (4/3) 3/2 and three (one of them unstable) for smaller values of h.
Calculating the quasistatic evolution of the system in the potential
(Eq. (11.37 by solving (Eq. (11. 11 one may find that the region of the
pinning field (or potential) corresponding to I x I < 1j.J3 is never reached.
Consequently, there is a certain freedom allowing to change the pinning field
without changes in the quasistatic evolution of the system. Precisely, one may
replace the pinning field on the interval - 1/.J3 < x < 1/.J3 by any function
smaller by the absolute value than (4/3 )3/2. still leaving the quasistatic
evolution unchanged. In Fig. 11. 10b three examples of such modifications are
shown.
The introduced modifications to the pinning field significantly change the
potential of the system at zero field. see Fig. 11. 1Oa. Imagine we brought the
system to the state (at h = 0), where it occupies the right energy well. This
536 K. L. Metlov

6
I
5 I
I
4 I
::: 3
I
I
~ 2 I
w I
I
/
0 /'~ /
/' "- /
-1 /' "-
-2 -1 0 2
Coordinate, x
(a)
2
(4/3)3/2
.......................... ,..
. .,....---...,
\ I
\
~ 1 \
~ \

""~ 0
0/) II\
\

'"
\
\
i:>: -1 \

j.4!3.!:12 f----<][]-\-"-!"

-2 '-----_----'---'----'-_--'-_--'-----'--_ _---1
-2 -I -1/3 1/ 2 0 1/3 112 I 2
Coordinate, x
(b)

Figure 11. 10 The quasistatic evolution under the applied field of systems described by all
potentials shown in (a) is exactly the same. This fact is clear from corresponding pinning
fields shown in (b). Arrows in (b) are drawn to guide the eyes. The dashed line on both
figures shows the original potential (Eq. (11.37 and the corresponding pinning field.

can be achieved by decreasing the external field from the starting value above
(4/3)3/2. At a low enough temperature the kinetics of the thermal relaxation
from this initial state are dramatically different in all three considered cases.
While in case I in Fig. 11. lOa the system will quickly pass to the left energy
well, in case III it will spend a considerable time in the right one. Thus, we
have just shown by a simple example that defining the quasistatic evolution of
the non-equilibrium system does not, generally speaking, define its relaxation
kinetics.
The quasistatic evolution of the system in the potential (Eq. ( 11 . 37
serves for the illustration of the basic idea of incompleteness of information
contained in the quasistatic response, but evolution of such system cannot
berepresented by the PM. Consequently, it can not be directly translated to
the systems representable by the PM. Of course, the quasistatically
Preisach Model and Simulation of Relaxation Kinetics 537

unreachable parts on the energy landscape are present in any system with
hysteresis. However, the PM is usually applicable to systems involving much
more randomness a Eq. (11. 37). It could happen that the differences in
unreachable parts are averaged-out, or reachable, and unreachable parts of
the potential are statistically similar to each other, so that little information is
lost.
The family of periodic potentials with random phase considered in the last
section allows illustrating this fact and determining the uncertainty related to
the unreachable parts of the potential. Let us turn back to the potential
(Eq. (11.33, so that
U(X) =- UoCOS[21T(XI0 + r)]/(21T) + kX(GX - H) (11.38)
where the notation of the previous section is used and U o = H co 0, k is the
coefficient relating the magnetization and the average domain wall
displacement, for parallel stripe domain structure k =2MoIO s ' where M o is
the saturation magnetization of the material and Os is the period of stripes.
To model the relaxation kinetics each separate system of the ensemble
was first subjected to the field history H = 00-0 (which process was supposed
to be quasistatic and at zero temperature) , and then instantly connected to the
thermostat having the finite temperature T. The coordinate of the system after
the application of the above field history was found according to the procedure
described by the rules (Eq. (11. 13. After this procedure the system is
always found in some local energy minimum. The relaxation after the
connection with a thermostat is established (supposing the dynamics are
overdamped) is governed by Smoluchowski equation

a [U' (x)
2
a ] a
aTP(X,T) + ax ----u;;---P(X,T) - v ax 2P (X,T) =0 (11.39)

where x = X lOis the reduced coordinate, T = Uo t I ( y02) is the reduced


time, P (x, T) is the probabil ity distribution function of the system coordinates
at a time T, Y is the damping factor, v = k BTl U o is the reduced
temperature, k B is Boltzmann constant, prime means the derivative of the
function with respect to its expl icitly given argument. During the simulation the
coordinate of each system was first averaged over the thermal fluctuations
<x ( T = fx P (x, T) dx and, then, over the ensemble to obtain <x ( T .
The simulations for the potential (Eq. (11.38 and its two modifications
(shown by the dashed and dotted lines in Fig. 11. 11) were performed with
v = 3/2 and 2M s 0 2 G I (U o0 s) = 1/8. The corresponding zero field relaxation
curves are shown in Fig. 11.12. At small times these curves follow logarithmic
law very closely, at large ones they are exponential.
The considered modifications of the potential correspond to the extreme
cases of maximum (positive or negative) displacement of the global energy
minimum (Fig. 11. 11) sti II leaving the quasistatic evolution in tact. Consequently,
538 K. L. Metlov

\
\
6 \
\
\
, :--
4
5 '-, \

4 " , .......
C , ",'

3 ..... b " ... _...... / ....~


"0
vtC ."
:-:,..
.... _.... ....---",

b
o
a
-1
c

-3 -2 -I 0 1 2 3
Coordinate, x

Figure II. 11 A single realization of three considered choices of random potentials,


different types of lines show the original potential Eq. (11. 39): (a) and its
modifications (b), (c) which respond to the external field exactly in the same way.

----------------

-[

-., .
" . ....
-2L-.------"-----------'-----'--------'
o 5 10 15 20
Time, r

Figure 11.12 Time profile of relaxation, the solid curve is the one for original system
Eq. (11.38), two other curves correspond to the modifications of the potential shown in
Fig. 11. 11.

for other modifications corresponding to the fixed quasistatic response, the


relaxation curve may lie anywhere in between of the bounds shown in Fig. 11.12.
This shows that the relaxation kinetics of a non-equilibrium system with
non-local memory are generally speaking, not defined by its quasistatic
response. There are relaxation models (Bertotti , 1996; Mayergoyz and
Korman, 1991, 1994), which describe the system in terms of the Preisach
Preisach Model and Simulation of Relaxation Kinetics 539

distribution function and give a unique answer for the relaxation curve at a
given field history. According to the above discussion, such models involve
certain assumptions about the unreachable parts of the potential allowing us to
select that unique relaxation curve from an infinite set of curves located in
between of the bounds similar to ones shown in Fig. 11. 12. These assumptions
may need clarification in the future.

11. 9 Summary

In this chapter several examples of hysteresis non-linearities were reviewed


and representations of their response operators by the Preisach model were
built. There is a wide array of systems that are exactly representable by the
PM, but not all of the systems with hysteresis are. Certain self-consistency
requirements for the system to be represented by the PM were formulated in
terms of simple to check in an experiment necessary conditions (Eq. ( 11 . 8))
for the representability of the evolution of a particular system by the PM.
It is possible to formulate the sufficient conditions of the representability of
the average evolution of ensemble consisting of systems in a random one-
dimensional potential by the PM. Such evolution is representable if the second
derivative of the potential is a stationary random process.
Several calculations of the Preisach representations of the model systems
obeying such a sufficient condition were presented here with the emphasis of
tracking the dependence of their representation on the force that stabilizes
each of the systems in the ensemble. The resulting dependencies are in
qualitative agreement with the experiments and the earlier works using a less
detailed description of the system's response.
The completeness of the information stored in the PDF from the point of
view of predicting the relaxation kinetics of the system was analyzed on the
example of the evolution of ensembles built around the periodic potentials with
a random phase. It turns out that systems having the same Preisach
distribution function, which exactly describes their rate-independent
hysteresis, may have the relaxation varying within considerable large limits.
Thus, the information stored in the Preisach distribution is incomplete for
obtaining a unique answer on how the system would relax. The choice of the
unique relaxation curve has to be made on additional assumptions beyond the
framework of the Preisach model.
540 K. L. Metlov

References
Bertotti, G. Phys. Rev. Lett. 76(10): 1739 (1996)
Bertotti, G., V. Basso and G. Durin. J. Appl. Phys. 79: 5764 (1996)
Bertotti, G., Hysteresis in Magnetism. Academic Press, San Diego (1998)
Bertotti, G., I. D. Mayergoyz, V. Basso and A. Magni. Phys. Rev. E 60
(2): 1428 (1999)
Friedman G., L. Liu and J. S. Kouvel. J. Appl. Phys. 75(10): 5683 (1994)
Jiles,D. C. and D. L. Atherton. J. Magn. Magn. Mat. 61: 48 (1986)
Koey, C. and U. Enz. Philips Res. Repts. 15: 7 (1960)
Korn, G. and T. Korn. Mathematical Handbook, 2nd edn. McGraw-Hili Book
company, New York (1968)
Mayergoyz, I. D. Phys. Rev. Lett. 56, 1518 (1986)
Mayergoyz, I. D. Mathematical Models of Hysteresis. Springer-Verlag,
Berlin (1991)
Mayergoyz, I. D. and C. E. Korman. J. Appl. Phys. 69 (4): 2128 (1991)
Mayergoyz, I. D. and C. E. Korman. J. Appl. Phys. 75 (10): 5478 (1994)
Metlov, K. L., Jana Kadlecova and Ivan Tomas. J. Magn. Magn. Mat. 196-
197: 196 (1999)
Metlov, K. L. Physica B 275: 164(2000)
Oti, J., F. Vajda and E. Della-Torre. J. Appl. Phys. 69(8): 4826 (1991)
Pust, L., G. Bertotti, Ivan Tomas and G. Vertesy. Phys. Rev. B 54(1): 12,
262 (1996)
J. P. Sethna, K. Dahmen, S. Kartha, J. A. Krumhansl, B. W. Roberts and
J. D. Shore. Phys. Rev. Lett. 70: 3347-3350 (1993)
Tomas, I. J. Magn. Magn. Mat. 87: 5 (1990)
Tomas, I., G. Vertesy, K. L. Metlov. J. Magn. Magn. Mat. 150(1): 75
(1995)

The financial support of the Grant Agency of the Czech Republic under the projects 202/99/
P052 and 101/99/1662 is appreciated. I would like to thank Ivan Tomas for his reading of the
manuscript and many valuable discussions.
12 Antiferromagnetism of Mn Alloys

Akimasa Sakuma, Kazuaki Fukamichi

12. 1 Introduction

Most Mn alloys exhibit antiferromagneto properties due to their occupation of


3d orbitals close to half-filling. There are many systems having a complex
(non-collinear) magnetic structure, depending on the number of d-electrons
and the crystal structure associated with the geometrical frustration of
moments. Experimental evaluation of these magnetic structures is quite
difficult, and hence there still remain the systems whose magnetic structures
have not yet been establ ished. Practically, on the other hand, the appearance
of the giant magnetoresistance (GMR) and tunnel magnetoresistance (TMR)
are very important for electronic and magnetic devices such as magnetic
recording heads and magnetic random access memory (MRAM) cells. In these
devices, the antiferromagnetic (AFM) Mn alloy systems are useful as an
exchange biasing film for the ferromagnetic (FM) film. Therefore, microscopic
foundations of the AFM Mn alloys are strongly desired to develop excellent
properties for GMR and TMR.
In this chapter, we look over, based on our recent works, physical
aspects of antiferromagnetism of Mn alloys both from theoretical and
experimental viewpoints in order to provide a base for the future development.
The construction of this chapter is as follows. In the next section, we give
microscopic insight into the itinerant-electron magnetism using the Hubbard
model in order to consider the condition of the magnetic ordering including
antiferromagnetism, and derive a useful equation for the magnetic excitation
energies from the magnetic ordered states. In Section 12. 3, the theoretical
frame discussed above is applied to the first principles version to deal with the
actual Mn alloy systems. In Section 12. 4, based on the first principles
approach described in the third section, we present the calculated results of
electronic and magnetic structures of some Mn alloys. Details of the
experimental data of antiferromagnetism of Mn alloys will be given in
Section 12.5, and the AFM properties related to magnetic and electronic
devices are discussed in Section 12.6. Finally, concluding remarks are given
in the last Section 12.7.
542 Akimasa Sakuma, Kazuaki Fukamichi

12. 2 Theory of Itinerant Electron Magnetism with the


Hubbard Model

12.2. 1 Model for the Itinerant Electron System


Let us first start with the single band Hubbard model (1963, 1964) given
below:

H =- ~ ~tijC-:-aCja + U ~nil nil (12.1)


a

where Cia ( C ~) denotes the electron annihilation (creation) operator of (J (= t


or ~) spin at the i-th site and n ia C ~c ia is the number operator. As shown in
Fig. 12. 1, the first term represents the electron hopping between the i-th and
the j-th sites with probability amplitude of t ij , the second term expresses the
Coulomb interaction with energy U between electrons of opposite spins
encountered at the same site. Though, boldly dropped multi-orbital and intra-
atomic Coulomb interactions between them play an important role especially
for ferromagnetism, we restrict ourselves to this simple model to avoid the
confusion from plural effects. Even though it is simple, the model can hardly be
solved exactly except for a few cases, because the second term reflecting
particle character of the electrons puts the treatment into typical many-body
problem.
u

Figure 12. 1 Schematic picture of the Hubbard model. The circles represent the atomic
positions and the arrows represent the spin direction of electrons. The electrons can hop to
another site with a probability amplitude of t and interact with another electron with the
opposite spin direction when they encounter at the same site.

12.2.2 Path-Integral Approach for the Itinerant-Electron


Magnetism
A powerful tool to describe an ordered state and the (thermal or quantum)
fluctuations in the many-body systems is the path-integral approach. In this
method, using the Hubbard-Stratonovich (H-S) transformation for the second
term in Eq. (12. 1), the partition function can be expressed as (Hubbard,
Antiferromagnetism of Mn Alloys 543

1959; Wang et al., 1969; Korenman et al., 1977; Schulz, 1990)

Z = fO~OTJexp(- S{~'TJ}) (12.2)

where the integral fox means fox = J~~ IT r;r (U / 41T) fdx 1/2 i ( Tn) with Tn =
n/3 / N (/3 = 1/ k B T). The action S is given by

S{~, TJ} =- Trln{oT - t "" u"


+ E} + 4"/3 L.J (~~ + TJD (12.3)
I

where A designates the matrix A= II A ij II and the matrix elements of E is


given by

(12.4)

In Eq. (12.3), Tr denotes trace over site (i), spin (J) and time ( T), and (1

in Eq. (12.4) is the Pauli matrix. Here, we use a static approximation for the
auxiliary fields {~, TJ}, regarding the fact that the spin excitation energy is a
several tens of meV at most, while the band energy of the electrons is of the
order of eV. As shown in Fig. 12. 2, the virtue of the H-S transformation is a
transformation from a many-body problem to a one-electron model under the
potential given by Eq. (12.4).

o o 0
I~
~JAuxiliary field) ==:>
[:1 Bi (Exchange field)
SP approx. =U/2 (m;)

Figure 12. 2 The Hubbard-Stratonovich (H-S) transformation in the path-integral


approach makes the Hubbard model to be a single-particle problem under the auxiliary
field ~" In the saddle point (SP) approximation, ~i is fixed at the exchange (molecular)
field B,.
544 Akimasa Sakuma, Kazuaki Fukamichi

12.2.3 Saddle Point (Molecular Field) Approximation

A magnetic ordered state in the Hubbard model, if it exists, corresponds to a


saddle point (SP) of the action S given in Eq. (12.3). The SPs turn out to be
a molecular field solution which can be given by a usual self-consistent band
calculation. As will be discussed later, a slight deviation of the moment
direction at the given site leads to the change of the action from the SP, which
can be treated within the single-site multiple scattering theory. The
Heisenberg-type exchange constant can be derived from the energy change
due to this deviation of the action. This enables us to treat the exchange
constant within the local density functional approximation (LOA) which will be
presented in Section 12. 3. 3.
In Fig. 12.3, the $P of the action S in Eq. (12.3) is obtained from
(()S/dT)i =()S/d~i =0, resulting in

T)i = i ~Tro'T(-()T+t-E)iil = i<n i >,


_ 1 ' '-1 _
~i - 73Tro.T(- ()T + t - E)jj 0' - <mi>'

where <n i > and <m i > are the expectation values of the electron number and
the magnetic moment, respectively, at the i-th site. At the SP, the action is
written as

S5 =-Trln{- G- 1 ) + ~,B~[<mY - <nY] (12.5)


I

where G is the Green function (GF) given by

(m)
i( n)

Figure 12.3 Schematic picture of the action as a function of auxiliary fields ~ and 1). The
SPs with respect to ~ and 1) correspond to the expectation values of the magnetic moment
<m) and the electron number i <n), respectively.
Antiferromagnetism of Mn Alloys 545

[G- 1 Jij =-()JJ ij + t;j - (~<ni) - ~<mi)' a)Oij ' (12.6)

The action in this case acts as a thermodynamic potential and the effective
Hamiltonian (EH) can be expressed as

H MF =- ~ ~t;jci;,Cja + ~vin; - ~Bi' mi + ~ ~m;)2 - <n;)2)


IJ (1 I I ,

(12.7)

where V; = ~<n;),Bi = ~<mi),m; = 2:ci;,(a)aaCia,ni


a,a
which <n;) and <m;> are respectively given by

en;) = <nil) + <nil)'


em;) = m;(cos;sin8; ,sin;sin8; ,cos8;)

with
m; = (afx + afy + afz) 1/2 ,

a;x = 2Re<ci t ci, ),aiY = 2lm<cit ci, ),a;z = <nil) - <nil)'


8; = cos-1(aiX/m;), ; = tan- 1(aida;x)'
It can easily be understood that Eq. (12,7) is nothing but the molecular field
Hamiltonian derived from the following approximation for the second term in
Eq. (12.1):

ntn,-<nt)n, +<n,)nt -<CtC,)c+Ct -<cjc )CtC


- <nt )<n,) + <Ct c, )<cj Ct)

Here, < nil)' < n; ), < cit C iI ) and < Ci C iI ) should be solved self-
consistently, from which the amplitude m; and the direction (8 i , ;) of the
magnetic moment at the i-th site can be obtained. The vector B i in the third
term in Eq. (12. 7) corresponds to the exchange (molecular) field acting on
the electron at the i -th site, proportional to the magnetic moment <m;> brought
about by itself (Fig, 12.2). The last term in Eq. (12.7) is the correction term
for the total energy due to the double counting of the interaction energy in the
molecular field approximation (MFA).

12. 2. 4 Rotation of the Local Spin Axes in the Complex


Magnetic Structures
In principle, the direct diagonalization of Eq. (12.7) can give any magnetic
structure. In the practical calculations, however, it is useful to rotate the spin
quantization axis to the direction of the exchange (molecular) field B; at each
position as shown in Fig. 12.4. This procedure enables us to avoid treating
the off-diagonal elements of the exchange potential in the spin space, which is
unadvisable in the conventional LOA scheme. Mathematically, this can be
546 Akimasa Sakuma, Kazuaki Fukamichi

performed by the transformation of the basis operators as

:' ~ yf --e-- \ --e-


z z

~YrY~~ z z
Figure 12.4 Rotation of the local spin axis to the direction of the exchange (molecular)
field on each site.

(CiI) =
Cil
U(e u 4>J (d
dil
il
), (12.8)

where U is the rotation matrix of spin 112 given by


_ (e-i;l2cose,12 - e-i;l2sine,/2)
U (e, , 4>,) - . /2' 12 . /2 1 (12.9)
e'~i sme, e'~i cose, 2

By the rotation, the EH can be rewritten as

RMF =- ~ ~ f'ij"dtd ju' + ~ ~v,dtdju - ~ ~ m,(djt d il - dj j d,j).


1/ ua I (] /

(12.10)
One can see that the exchange potential is expressed along each spin
quantization axis with no off-diagonal elements. Taking the place of the
exchange potential, the hopping integrals take charge of the non-collinearity of
the moments through the following transformation:

(12.11)

which connects the local spin axis between the i-th and j-th sites. In
Eq. (12. 11), the sign + is taken for a = a' , and - for a"#- a'. The relative
angle e'j between the moments at the i-th and j-th sites is defined as
COSe'j = cose,cose j + sine,sinejcos(4), - 4>j)
and aft is the phase factor which is efficient when 4>, varies with the site.
The quantities of m, and (e" 4>,) in Eq. (12. 10) should be solved self-
consistently by using
m, = TrP,tJ' (12.12)
where T r means the trace over spin and P is the density matrix defined in the
global frame of the spin space. Generally, this can be obtained from GF of
Eq. (12.6) as
Antiferromagnetism of Mn Alloys 547

Pi = ~ ~G;;(jWn) = 2~fdEf(E)[G;;(E + i5) - G;;(E - i5)J

(12.13)
with
W n = (2n + 1)//3
where f ( E) is the Fermi distribution function and 5 is an infinitesimal. GF
G (w) in the above equation can be given through the following
transformation:

G;; ( w) = U ( 8 i , 4>;) G;; ( W ) U+ (8;, 4> ;) ( 12. 14)


where GF G(w) is defined under the Hamiltonian of Eq. (12.10) by
G(w) = (w - HMF ) - I . (12.15)

Once the density matrix is obtained, m i and (8 i , 4> i) are given expl icitly as:
8i = cos- [(pi t -
1
pi t )/miJ,
4>; = tan- (Impi t
1
/Repi t ),
m; =1 mi I. (12.16)
When these quantities coincide with the input values in Eq. (12. 10), one can
reach the self-consistent results.

12. 2. 5 Magnetic Excitation Energy and the Exchange Constant


We provide a method to calculate the magnetic excitation energy in the
magnetically ordered states including non-collinear structures which can be
obtained from the SP approximation. Roughly speaking, the strategy is to
evaluate the energy change due to a slight deviation of the direction of a
moment at the ground state (SP). By using this approach, one can derive the
effective exchange constant acting on a given moment, in combination with a
generalized molecular field theory by Liechtenstein et al. (1985). The
magnetic transition temperatures, the Neel temperature in the present case,
can be estimated from this value.
According to the local force theorem, in Fig. 12. 5, the variation of the
action due to a slight change of the moment direction at the O-th site can be
given by

5S = - /3 ~:o 1mf dEf (E) { [G 60 (E) - G M(E) J


+ Um 0 G 66 (E) G ot (E) } [1 - cos ( 58) ]. (12.17)

Here, we have used the identity Trln A = In det A and the approximation,
Trln( 1- A)= - A, and exploited the fact that GF' s G are diagonal with
respect to the spin space in each local spin axis. The quantity 5S / /3 can be
548 Akimasa Sakuma, Kazuaki Fukamichi

~;zf?~~

/~~~
O-site

~I?
z~

L X

Figure 12.5 Magnetic excitation due to a slight deviation of the moment direction at the 0-
th site.

regarded as a variation of the thermodynamic potential due to a change of the


moment direction (88) at the O-th site. One can notice, by comparing with the
classical Heisenberg model H =- .:8J ije i e i (e: the unit vector), that the
i.j
coefficient of (1 - cos( 88 in Eq. (12. 17) can be related to the total sum of
the exchange fields acting on the moment at the O-th site ((.:8 J iD ei) eo).
i*O

Defining this coefficient by J o (= (.:8 J iDe;) eo), we obtain the following


i*O
expression as an effective exchange constant:

J o -- - Urn f It
4rr0 1m dEf(f){ Goo [- - II
(E) - Goo (E) ] + UmOG- OOIt - 'I
(E)G OO (E)}.

(12. 18)
This is the same form as that obtained for the collinear case by Liechtenstein
et al. (1987). Note, however, that GF used here is defined in the local frame
of the spin space rotated to each direction of the moment (Sakuma, 2000).
Especially, in the collinear case, using the sum rule for GF as

(G It - G tb) =- .:8G 61
j
(Urn j) G n
one can obtain the pair exchange constant between the i-th and j-th sites as

(12. 19)

Here, let us make a brief inspection on the effective exchange constant given
Antiferromagnetism of Mn Alloys 549

by Eq. (12. 18). The virtue of the quantity J o rather than J ij is available for
the estimation of the magnetic transition temperatures within the framework of
the MFA. In the three-dimensional spin systems, the magnetic critical
temperatures can be given by 2J o /3k B in the MFA scheme.
Figure 12.6 shows the calculated results of J o as a function of the electron
density n for the semi-elliptic density of states (DOS) (Sakuma, 1999). The
local magnetic moment m in Eq. (12. 18) is derived self-consistently from
mo=-(l/rr) 1m fdEf(E)(G66 (E)-GM (E. Finite mo starts at around
n = 0.25. One can easily find that this is the Stoner critical point for the
existence of ferromagnetism; i. e., PF U = 1, where PF is the DOS at the Fermi
level E F Although the moment mo increases with increasing n to mo = n = 1,
the sign of J o turns to negative at around n = 0.65, and then makes the FM
state unstable in the range O. 65 ~ n ~ 1. 35. This is qual itatively ascribed to
the Pauli principle which restricts the electron hopping in the FM configuration.
It is seen that the increase of U raises J o toward the positive side and
stabilizes the FM state. The position of n giving the maximum J o is unchanged,
remaining at n "" 1. 5. We emphasise here that the variation of the electron
density (number of electrons per orbital) in the right half region (1~n~2)
corresponds to the series of 3d transition metals from Cr or Mn to Cu metals.
Although the electronic structures of these materials are so intricate and
different from each other, the curves in Fig. 12.6 obtained from a quite simple
model with varying only the electron density reproduce well the relative
difference among the magnetic properties of these metals. Accepting this
premise, we can infer the following from Fig. 12. 6. The electron densities of
d orbital of Mn and Fe metals may be located around or across the point for
J o = 0, so the magnetic structures (ferro- or antiferromagnetism) of these

0.10 1.2

U=2D
1.0
0.05
~

0.8 co
2,
0
E
0.6 "
e E
0
E
~ -D.OS ";;j
()
0.4 0
....l

-D.10
0.2

-D. 15 '--------'--_---'-_ _-----J..._ _--''--~--'


o 0.5 1.0 1.5 2.0
Number of electrons (l/orbital)
Figure 12. 6 Local magnetic moment ma and the effective exchange constant J a
calculated from Eq. (12. 18) as a function of the number of electrons per orbital. The inset
shows the assumed DOS (Sakuma, 1999).
550 Akimasa Sakuma, Kazuaki Fukamichi

metals strongly depend on the circumstances such as the crystal structure, the
lattice constants, alloying and so on. On the other hand, the electron density
of Co may be located around the stationary maximum point. Therefore, the
value of J o is stable against the circumstances and it may maintain a high Curie
temperature T c . To see this more quantitatively, we have prepared the first
principles approach for actual Mn alloy systems in the next section.

12.3 First Principles Approach for the Magnetic


Structures of Transition Metal Systems

In this section, the theoretical framework discussed above will be mapped into
the linear muffin-tin orbital (LMTO) (Andersen, 1975; Skriver, 1984) method
based upon the density functional theory in order to make a quantitative
evaluation of the magnetic structures and their stability in the actual magnetic
systems including substitutional disordered alloys. For the disordered alloys,
we take the coherent potential approximation (CPA) (Soven, 1976; Taylor,
1967; Velicky et aI., 1968) within the framework of the LMTO method. It is
worth noting that the treatment given below can also be applied to the
Korringa-Kohn-Rostoker (KKR) method in a similar way.

12.3.1 Tight-binding (TB)-LMTO Method for Complex


Magnetic Structures
As mentioned in Section 12. 2. 3, the Hamiltonian for the spin polarized band
calculations has the same form as Eq. (12. 7). This equation has a
correspondence with the LDA Hamiltonian through the following relations:

- t - - VZ + VHartree + V nuclei'
1
Vi - 2[V i l (r) + Vii (r)],

1
Bi -- 2[Vil (r) - Vii (r)]ei.

Here, the term VHartree is the Hartree potential and V nuclei the potential from the
nuclei. The potential V Ie (r) which depends on the spin (J is given by the
functional derivative of the exchange correlation energy E xc with respect to the
electron density n ie (r) as

oExc
( r) = --,---------'-;"...,- (12.20)
Vie
on ie (r) .

The correction term, corresponding to the last term in Eq. (12. 7) for double
Antiferromagnetism of Mn Alloys 551

counting of the interaction energies, is given by

E de = -"21 ff '
drdr
n(r)n(r')
I r _ r' I - f[ dr n t (r) v t (r) +nI (r) v I (r) - E xc]

(12.21)
with nCr) = ~nu(r) and nd(r) = ~n; (r) . The resultant form of the LOA

Hamiltonian can be given as

(12.22)

with

(12.23)

and

(12.24)

where e; = (cos 11'; sin 8 p sin 'Pi sin 8pcos 8 i ) indicates the direction of the
molecular (exchange) field acting on a moment at the i-th site. Although, the
direction of the molecular field generally varies as a function of position r, we
adopt the rigid spin approximation, assuming that the direction is constant
within each muffin-tin (MT) sphere. When each moment points to the different
direction on the individual site, it is convenient to rotate each spin axis within
the MT sphere to the direction of the molecular field.
The Green function (GF) in the TB-LMTO scheme is given by
G(w) = Ll-1/2gY(w)Ll-1/2 (12.25)

with

(12.26)
where 9 Y ( w) is the so-called auxi Iiary GF constituted of the potential function
p Y (w) and the screened structure constants S Y respectively given by

pY (w)
w-C
= -Ll--fh.,J>u"u (12.27)
$Y = S (1 - yS) -I . (12.28)
Here, L =( i , I ,m) (i denotes the site, I and m are the orbital indices), and
Ll , y, and C are the LMTO-parameters determined by using LOA within each
atomic sphere. For the non-collinear spin structure, the structure constants S
should be expressed, in the same way as in Eq. (12, 11), by

SL'U',LU = (Uj-SL',LU;)u',u (12.29)


where S is the "bare" structure constant and the matrices U are defined by
552 Akimasa Sakuma, Kazuaki Fukamichi

Eq. (12.9).
Since at this stage the site-oft-diagonal elements in the screened structure
constants depend on the species of atoms through the values of y, further
steps are required for CPA. To the end of this sub-section, we follow a
proposal by Kudrnovsky and Drchal (1990), where the auxiliary GF of y
representation is transformed into the most localized ({3) representation
through the following relation val id for arbitrary representation:
gY = (pY)-l _ (pY)-lp~(pY)-l + (pY)-lp~g~p~(pY)-l (12.30)

where the matrices in the (3 representation are given by the following


equations:
g~(w) = [p~(w) - $73]-1, (12.31)
p~ = pY[l - ({3 - y)pYJ-I, (12.32)
? = $7[1 - ({3 - y) $7J- 1
= $(1 - {3$)-1. (12.33)
For closed-packed lattices, it has been found that the optimum values of {3 are
given by {3,=O = 0.3485, {3,= 1 = 0.0530, (3,=2 = 0.0107 which give the fastest
and essentially monotonic decay of ? in the real space. It is of importance
that? becomes free from the species of atoms.

12.3.2 Coherent Potential Approximation (CPA) for


Disordered Alloys
For non-collinear magnetic structures in the substitutional disordered alloys
having several sub-lattices, we apply CPA individually in each sub-lattice. The
CPA condition in the i-th sub-lattice is given by

~i(W) = ~ciagfa(W) (12.34)


a

or equivalently

<T;> = ~C;a[pfa(W)-Pi(w)J{l+[pfa(W)-Pi(w)J~;(w)}-1 =0
a
(12.35)
where Cia denotes a density (probability) of a atom occupying the i-th sub-
lattice. The quantity ~ i ( w) is the coherent propagator of the i -th sub-lattice,
and given by

(12.36)

where Pi ( w) and $ ~ are the coherent potential function and the screened
structure constant, respectively, in the {3 representation in the k-space. The
quantity gfa (w) is the local (auxiliary) GF of a atom in the i-th sub-lattice
Antiferromagnetism of Mn Alloys 553

which is given by q, i (w) based on the multiple scattering theory as


gfa (w) = q, i (w){ 1 + [pfa (w) - Pi (W)] q, i (W) } -1 ( 12. 37)

where p~ ( w) is the potential function in the {3 representation. In Eq. (12.37),


pfa (w) is diagonal with respect not only to (f, m) but also to spin a, since it
is defined in the locally rotated spin axis. Note, however, that the quantities
q, ; (w), gfa ( w) and P; ( w) generally have off-diagonal elements with
respect to (I, m) and a. Once gfa ( w) is solved self-consistently, GF G,
defined in the local frame, can be obtained by gia (w) through the
transformation of g~(w) to gia(w) with Eq. (12.30). The concrete form is

= (YM-{3M) OMM'
.11 M + (YM - {3M)(W - OM)
.11 1/2
+ .11 M + (YM - ;M)(W - OM) [g~ (W)]MM'
fl~2
x---.,...---7'-.,...-.,...-----,- (12.38)
.11 M, + (YM - {3M')(W - OM')

where M = ( I , m, a). G is further transformed into G defined in the global


frame through Eq. (12. 14) to get the density matrices. Calculating m ia' the
moment of a atom at the i -th sub-lattice by Eqs. (12. 12) and (12. 13), one
can obtain the average moment at the i-th sub-lattice by mi - ~Ciamia'
a
When the direction of this moment coincides with the input angle (e;, <Pi) the
self-consistent solution is achieved through S in Eq. (12.29).

12. 3. 3 Effective Exchange Constant


Based upon the scenario discussed in Section 12. 2. 5, the effective exchange
constant can be calculated in the LOA scheme. As mentioned in Section 12. 2.
5, the magnetic transition temperatures can be estimated to be 2J o/3k a by
the values obtained within the MFA scheme.
In the expressions of J o and Jij given by Eqs. (12. 18) and (12.19),
respectively, the matrices (Um) G ( w) can be rewritten in the TB-LMTO
method as

with O(w)-Pt (w) - PI (w). Then, we have expressions for J o and J ij in


the LOA scheme as

Jo =- 1
41Tlm f
dEf(E)Tr,m{Oo(E)[gY II II
oo (E) - gY oo (E)]

+ 0 0 (E) 9 Y ~~ (E) 0 0 (E) 9 Y ~~ (E) } ( 12.39)

and
554 Akimasa Sakuma, Kazuaki Fukamichi

J ij 1 f
= 41Tlm ij (E)Qj(E) gYji
dEf(E)Tr,m{Q,(E) gY II II (E)} (12.40)

where Trim represents the trace over the orbital part (I, m). Note here again
that J o of Eq. (12.39) can work both for collinear and non-collinear magnetic
structures, while J ij of Eq. (12.40) can be used only for the collinear case.
In disordered alloys, G should be calculated in the CPA process by using
the form of Eq. (12. 38). When the number of magnetic components in the
alloying system is greater than two, the Neel temperature TN is considered to
be determined with the largest value of J o , because the magnetism of the
system should be dominated by the component with the largest J at around the
critical temperature.

12.4 Electronic and Magnetic Structures of y-Mn


and Mn alloys

In this section, we show the calculated results of magnetic structures of several


Mn alloys based on the first principles approach discussed above.

12.4.1 y-Mn

As mentioned in the introduction, almost all Mn alloys exhibit


antiferromagnetism due to their occupation of 3d orbitals close to half-filling.
Among them, a number of face centered cubic (fcc) alloys, such as FeMn and
Mn3 Pt, have a complex spin structure other than a collinear or simple helical
structure. Exper imentally, on the other hand, y -Mn is considered to be a
collinear antiferromagnet, though it also has a face centered structure. It
should be noted, however, that the pure y -Mn does not exist metallurgically
and the actual system contains Cu or C atoms of a few atomic percents,
accompanied by a lattice distortion of c / a -- O. 95 (Endoh and Ishikawa,
1971) .
Theoretically, so far, several attempts have been made to clarify the
magnetic structure of pure y-Mn. Asano and Yamashita (1971) first carried
out the band calculation for pure y-Mn, assuming the collinear AFM spin
arrangement. Cade and Young (1980) proposed the 30 state (non-coil inear)
as the ground state, and the result was followed by Hirai and Jo (1985) with a
Hartree-Fock calculation. Based on LOA approach for non-coli inear magnetism
developed by Kubler et al. (1988), more recently, Fuj ii et al. (1991) showed
that the most stable magnetic structure is a collinear AFM as proposed
experimentally by Meneghetti and Sidhu (1957). Unfortunately, all of them,
except Asano and Yamashita (1971), neglected the lattice distortion of pure
Antiferromagnetism of Mn Alloys 555

y-Mn for the calculations. In this sub-section, based on the LOA calculations
by Sakuma (1998a), we focus our attention mainly on the effect of the lattice
distortion on the magnetic structure of y-Mn. The exchange-correlation term is
adopted as the form of von Barth and' Hedin (1972) with the parameters given
by Janak (1978). The core charge density is used as the so-called frozen
core. For the valence states, the s-, p- and d-basis functions are adopted,
neglecting the spin-orbit interaction. The atomic sphere radius is settled at
a = 3. 75A for the fcc structure.
The spin structure examined here is shown in Fig. 12.7, illustrating a
magnetic primitive cell constituted of four atoms in the face-centered structure.
e
Note that the angles and are defined individually for each site. As shown in
e e
Fig. 12.8, with settled at 45, = 0, = COS-I (1/,;3) =54.7 degrees and
e = 90 correspond to the so-called 1Q, 3Q and 2Q spin density wave (SOW)
structures, respectively, which are generally referred to as the multiple-Q
SOW (MQSOW) structures.

Figure 12.7 Magnetic primitive cell of an fcc lattice considered in literatures. Multiple-Q
spin density wave (MQSDW) structures, , Q, 2Q and 3Q states given in Fig. '2.8 can be
realized by putting e = 0, 90 and 54.7', respectively.

10 20 30
Figure 12.8 Multiple-Q spin density wave (MQSDW) structures in an fcc lattice.

Figure 12.9 shows the e dependence of the total energy of y-Mn with =
556 Akimasa Sakuma, Kazuaki Fukamichi

45 at c / a = 1 (Sakuma, 1998a). Starting from 8 = 0, the energy reaches to


the minimum value at around 8 = 20. This state, however, is not a self-
consistent solution since the output angle (8 s ) of the magnetic moments
deviates from the input angle 8 by 58 = 8 s - 8 = 5, whereas at 8 = 0 we have
58 = O. With increasing 8, on the other hand, the energy increases and has a
maximum value at around 8 = cos - 1 ( 1/,)3) = 54 . r.
At this point, 58 vanishes
and the solution becomes self-consistent. Actually, 58 changes its sign from
minus to plus at this point, getting toward to zero again at 8 = 90. The results
reveal that the 30 SOW structure (8 = 54. r) is remarkably unstable, and
within the other structures the 20 structure (8 = 90) is more stable than the
10 structure (8=0) by 0.08 mRy/atom. The results for y-Fe are also shown
in the same figure for comparison. The lattice constant is chosen at 3. 6A. In

fcc-Mn

2Q 3Q IQ

o 20 40 60 80
BC)
Figure 12. 9 The e dependence of the electronic total energy of y-Mn and y-Fe
(Sakuma, 1998a).

contrast to y-Mn, the 30 structure (8 = 54. r) has the lowest energy among
the given magnetic structures. Note, however, that the clarification of the
magnetic ground state of y-Fe may require super cell structures. More
sophisticated analysis for y-Fe has been performed by Antropov et al.
(1996).
Figure 12. 10 shows the 8 dependence of the total energy of y-Mn for
several values of c / a under the constant cell volume (Sakuma, 1998a).
Obviously the 20 structure is stable at c / a = 1. 05. Across c / a = 1, the total
energy for the 20 structure goes up abruptly and that for the 10 structure is
getting lower with decreasing c / a from unity. It is worth noting that the self-
consistent solutions are restricted only in 8 = O and 8 = 90, except for
Antiferromagnetism of Mn Alloys 557

c I a = 1 where the state of e = 54. ris also self-consistent. Therefore, the


stable state for c I a < 1 can be regarded as the 1Q structure. Furthermore,
the value of the total energy of the 1Q structure decreases with decreasing c I
a. The total energy E and the magnetic moment M at the 1Q structure are
plotted against c I a in Fig. 12. 11 (Sakuma, 1998a). The 1Q structure is
found to have a minimum energy at c I a ......, O. 9. Although this value is
considerably smaller than that expected from experiments (cia""'" 0.95), both
the magnetic structure accompanied by the lattice distortion of c I a < 1 and the
moment of around 2. 3j.ls are in agreement with the experiments. From these
results, it can be concluded that the lattice distortion to c I a < 1 is strongly
responsible for the collinear structure in y-Mn and concurrently the system is
stabilized by this distortion.

Y-Mn

i=45'

o 20 40 60 80
on
Figure 12.10 The angle e dependence of the electronic total energy of y-Mn for several
values of the axial ratio c / a (Sakuma, 1998a).

An intuitive interpretation of the mechanism that the magnetic structure


favors 1Q for cia < 1 and 2Q for c I a> 1 may be ascribed to a change of the
magnitude of AFM exchange interactions between Mn moments depending on
the inter-atomic distance. That is, when cia < 1, since the atomic distances
between the inter-c-plane Mn atoms are shorter, the exchange interactions
between them are expected to be stronger than that for intra-c-plane Mn atoms
and vice versa. Actually, applying the Heisenberg model in Fig. 12. 12, one
can reproduce both the 1Q and 2Q structures by changing only the relative
strengths of J 12 and J 13 . In order to examine the validity of this feature, the
exchange constants are calculated by the first principles approach using Eq.
(12. 40). The results calculated for the collinear structure of e
= O' with
different ratios of cia are listed in Table 12.1 (Sakuma, 2001a).
558 Akimasa Sakuma, Kazuaki Fukamichi

2.6

Y-Mn
2.5

E0 M
2.4 3
"
~,
c.::
~
~
C
.-. "0
E
2? 2.3
"
c
w E
~

AF
0=0 2.2
>.
c.::
E
'n
6
2.1
a.80 0.85 0.90 0.95 1.00
cia
Figure 12. 11 Electronic total energy E and the magnetic moment M of y-Mn as a function
of the axial ratio c/o. The angle e is settled at 0 (Sakuma, 1998a).

Contradictory to the above-mentioned inference, no reversal of the strength


between J 12 and J 13 can be confirmed in the change from c / a 1 to c / a> 1, <
and both values at c / a = 0.95 are smaller than those at c / a = 1. Especially,
it is noticeable that J 13 remarkably decreases with decreasing c / a. These
results suggest that the pair exchange constants are not locally dominated by
the inter-atomic distance only, and the relative strength of J 12 to J 13 is not
responsible for determining the 10 and 20 structures.

!
!

!
!
!
I
I
I
I J 13
I
I
!
I
I
I
I

Figure 12.12 Pair exchange constants J 12 and J '3 between the nearest neighboring Mn
moments in the magnetic unit cell of y-Mn.
Antiferromagnetism of Mn Alloys 559

Table 12.1 Pair exchange constants (J12' J 13 ) between the nearest neighboring Mn atoms
and the effective exchange constant J o = ~ J;o (all in units of meV) for cia = O. 95, 1. 0
;;toO

and 1.05 of y-Mn (Sakuma, 2001a).


cia J 12 J 13 Jo
0.95 -18 12 54
1.0 -18 16 68
1.05 -19 17 64

The effective exchange constant J 0 calculated by Eq. (12. 39) is also


shown in Table 12.1 (Sakuma, 2001a). The Neel temperatures TN estimated
from these values are in the range of about 420 to 500 K, corresponding to the
experimentally estimated value of about 480 K (Endoh and Ishikawa, 1971).
However, it should be noted that J o for c / a = 0.95 is smaller than others even
though it is energetically most stable. This trend can also be seen in J 12 and
J 13. One can infer from these results that the most stable magnetic structure
does not always have a higher magnetic stiffness than others against the
transverse fluctuation of the moment, because the magnetic stiffness reflects
the curvature of the action around the SP, whereas the stable magnetic
structure is determined by the global minimum of the action.

12.4.2 FeMn Disordered Alloy

Although the magnetic structure of FeMn disordered alloy has been


investigated by several experimental techniques, it has not been clarified yet
because of scattered results, depending on the experimental method. The 1Q
structure in Fig. 12. 8 was proposed by Bianti et a!. (1987) from inelastic
neutron scattering. Non-collinear structures, 2Q and 3Q structures, were
predicted by Kennedy and Hicks (1987) from M6ssbauer transmission spectra
where both the models are indistinguishable from each other. The neutron
diffraction cannot also distinguish the 1Q structure from the 3Q one. Using
neutron diffraction combined with other techniques, Endoh and Ishikawa
(1971) constructed the magnetic phase diagram of FeMn alloy system in a
wide range of composition where FeMn alloys around the equiatomic
composition have the 3Q type structure. One of the present authors performed
the theoretical electronic structure calculations to investigate the magnetic
structure and its stability (Sakuma, 2000).
Figure 12. 13 shows the electronic total energy of FeMn alloy as a function
e
of defined in Fig. 12.7 (Sakuma, 2000). The calculations were performed
by the TB-LMTO method combined with CPA as mentioned in the Section
12.3. The lattice parameter is settled at a = 3. 63A. With increasing e, the
560 Akimasa Sakuma, Kazuaki Fukamichi

energy decreases to a minimum value at around = 54. e


(30 state) and goes r
up to the value at e = 90. In a similar way, as given in the preceding section,
e
only at = 0, 54. rand 90 the output (obtained) directions of the moments
just coincide with their input angles, which reveals that the self-consistent
solutions are real ized only at these angles. Consequently, the results
demonstrate that the magnetic structure of FeMn alloy is the 30 SDW state, in
accord with the experimental results given by Endoh and Ishikawa (1971).
However, the moment is different in magnitude between them. Namely, the
average moment is about 1. 8 /.Is in the present case, whereas Endoh and
Ishikawa reported it to be about 1/.Is. To examine whether the larger moment
can be ascribed to the single site approximation as CPA or not, the
comparison with the results for hypothetical FeMn ordered alloy (CuAu-1 type
structure with c / a = 1) has been made. Figures 12. 14 a, b show the DOS of
ordered and disordered FeMn alloys with the 30 structure, respectively
(Sakuma, 2000). Although a slight smearing due to the randomness is
confirmed in the DOS, the characteristic features in the ordered phase are
found to be kept in the disordered phase. Actually, the imaginary parts of the
p
coherent potential functions (1m (w)) reflecting the life-time of electrons due
to the scattering by the randomness are quite small, compared with those for
MnPt and MnNi alloys. In such a situation, the average magnetic moments of
both phases are quite close in each magnetic structure as given in Table 12. 2
(Sakuma, 2001b). This leads us to consider that the similar nature of Fe and
Mn atoms gives little change of the electronic structures in the change of the
atomic configuration and CPA is not responsible for the discrepancy in the
moment.

FeMn

E
0
~
>-.
~
>-.
5
.-.
~
"-l
~~
0

IQ 2Q 3Q

0 10 20 30 40 50 60 70 80 90
en
Figure 12. 13 The e dependence of the electronic total energy of FeMn disordered alloy
(Sakuma, 2000).
Antiferromagnetism of Mn Alloys 561

40
FeMn (O-3Q)

E0 20
1il
c:
'c..
'">, 0
e:
C/)
0 20
Cl

Mn
40
-0.8 -0.6 --0.4 --0.2 0 0.2
Energy (Ry)
(a)
40
FeMn (DO-3Q)
E0 20
1il
c:
'c..
'">, 0
e:
C/)
0 20
Cl
Mn
40
-0.8 --0.6 --0.4 --0.2 0 0.2
Energy (Ry)
(b)
40

FeMn (DO-PM)
E0 20
1il
c:
'c..
'">, 0
e:
C/)
0 20
Cl
Mn
40
-0.8 --0.6 --0.4 --0.2 0 0.2
Energy (Ry)
(c)

Figure 12. 14 Electronic density of states (DOS) of FeMn with (a) 30 structure in the
ordered atate; (b) 30 structure in the disordered state and (c) the paramagnetic ( ==
DLM) state in the disordered state (Sakuma, 2000).
562 Akimasa Sakuma, Kazuaki Fukamichi

Table 12. 2 Calculated results of the total energy difference (I::. E), the magnetic moment
(M) and the effective exchange constant (J o ) of each magnetic structure of disordered (DO)
and ordered(Q) FeMn alloy. The upper line of M and J o represent the results for Mn atom
and the lower line for Fe atom, respectively (Sakuma, 2001b)
I::.E(meV/atom) M(fJB) Jo(meV)
10 12.8 1.75 57
1. 56 22
DO 20 3.2 1. 83 75
1.67 51
30 0 1.88 81
1.72 63
PM 65 1. 27
1. 39
10 9.0 1. 60 16
1. 49 44
0 20 -19.7 1. 87 104
1.54 27
30 -17.2 1.92 90
1. 61 54
PM 205 1.54
1.48

Next, let us consider the paramagnetic (PM) state in the disordered


phase as an extreme limit both of the atomic (chemical) and magnetic
disorderings. The PM state in this case is regarded as the disordered local
moment (DLM) state calculated by CPA, in which the Mn moments pointing
upward (Mn ( t and downward (Mn ( t , and those of Fe moments (Fe
( t ), Fe ( t are randomly distributed in the fcc lattice. The DOS in
Fig. 12. 14c turns out to be remarkably different from that in the magnetically
ordered states in Figs. 12. 14 a, b. The energy in this state is as much as 65
meV/ atom higher than that of the 30 structure in the disordered phase.
Therefore, one can find that the magnetic randomness in the FeMn system has
a strong influence on the electronic state, as compared with the chemical
randomness. Note, however, that the magnetic moment in this state is
1.3 J..Is' still larger than that reported by Endoh and Ishikawa (1971). The
moment is nearly independent of the magnetic structures, and as a result, we
can confirm that the calculated moment may lie around 1. 8 J..Is even for any
other magnetic structures. In short, 1. 8 J..Is can be confirmed as a theoretical
prediction for the moment of FeMn alloy, though the origin of the difference
from the experimental result is not clear.
In what follows, we show the calculated effective exchange constant J o of
FeMn disordered alloy, which gives information on a magnetic stability around
the molecular field being the SP in the auxiliary field. For systematic
understanding, we plot the behavior of J o as a function of the Fermi level E F ,
Antiferromagnetism of Mn Alloys 563

the upper limit in the integral range of Eq. C12.39), in which fCE) is treated
as a step function. The resultant curves are shown in Figs. 12. 15a, b, c

150
FeMn (DO-IQ)
100

50
>
5" 0
~
-50

-100

-150
-0.6 -0.4 -0.2 o 0.2
Er(Ry)
(a)
150

100 FeMn (DO-2Q)

50
>
5" 0
~
-50

-100

-150
-0.6 -0.4 -0.2 o 0.2
Er(Ry)
(b)

150

100 FeMn (DO-3Q)

50
>"
5 0
~
-50

-100

-150
-0.6 -0.4 -0.2 o 0.2
Er(Ry)
(c)

Figure 12. 15 Effective exchange constant J o of FeMn in the disordered state as a function
of the Fermi level Er(band filling) for each magnetic structure (Sakuma, 2000).
564 Akimasa Sakuma, Kazuaki Fukamichi

(Sakuma, 2000). The actual E F is located at the origin of the abscissa and the
variation of E F corresponds to the change of the electron number in the rigid
band scheme. The positive value at the actual E F ( = 0) indicates that the
directions of the magnetic moments assumed in the band calculation are stable
against a slight deviation of the molecular (exchange) field directions. The
value of J ofor the 30 structure with the largest average moment is larger than
that for the other two structures. A noticeable point is that the difference of J o
between 10 and 30 is considerably larger than the total energy difference.
This implies that the activation energy of the 30 structure is much larger than
the energy difference from the local minimum in the 10 state. That is to say,
the moment on each site in the 30 structure is sufficiently stable around the
molecular field, or the SP. The Neel temperature TN estimated from J o for the
30 structure of about 600 K is in good agreement with the experimental value
of around 480 K (see Fig. 12.33).
Schulthess et al. (1999) also proposed the 30 structure by the KKR
method combined with CPA. Their average moment is 1.98 J.JB' higher than
our result. Akbar et al. (1998) performed the molecular dynamics simulation
for the spin structure of FeMn alloys on the basis of the functional integral
technique for the generalized Hubbard model. They found that the magnetic
structure of Feso Mnso is quite complex, other than multiple-O structures, and
the average moment of about 1J.JB is relatively close to the experimental value.
However, the results may be strongly dependent on the parameters used in the
Hamiltonian model. Recently, the first principles calculations for clusters have
been performed by Spisak and Hafner. (2000). Unfortunately, as yet, no
consentaneous results have been obtained.

12.4.3 L1 o-Type MnPt, MnNi and MnPd Alloys

MnPt, MnNi and MnPd ordered alloys form L 10 ( =CuAu-l) type structure and
have a collinear antiferromagnetism. They have quite a high Neel temperature
TN, about 1000 K for MnPt (Kren et aI., 1968), and 1100 K for MnNi (Pal et
al., 1968), 850 K for MnPd (Kasper and Kouvel, 1959), and hence are
expected to have a highly potential application to the exchange biasing film for
GMR and TMR devices. The representative spin structure (AF-I) of MnPt is
illstalled in Fig. 12.16. Although the ordered phase is favorable in these bulk
systems, the disordered phase can be easily realized by means of the thin film
deposition process. It has been considered that the disordered phase is
technologically undesirable since the exchange bias field does not work when
such a film is used for GMR and TMR devices. However, little investigation
has been made on the magnetism of these disordered alloys, not only in the
bulk state, but also in the fi 1m structure.
Figures 12. 17, 12. 18 and 12. 19 show the DOSs of Llo-type ordered
phase of MnPt (Sakuma, 2000), MnNi (Sakuma, 1998b) and MnPd (Umetsu
Antiferromagnetism of Mn Alloys 565

Mn

Figure 12. 16 Crystal structure and co-Ilinear magnetic structure AF-I of L 10 -type
ordered MnPl.

50
40
MnPt(Ll o)
30
20
8'0
1;j 10
.:: 0
'-
;>, 10
e::
20
[fj
0 30
Cl
40
50
--D. 8 --D.6 -0.4 -0.2 0.2
Energy (Ry)
(a)
40

MnPt (L1 a-PM)


8'0 20
1;j
c
'5-
en
>, 0
e::
[fj
0 20 Pt
Cl
Mn
40
-0.8 -0.6 -0.4 --D.2 o 0.2
Energy (Ry)
(b)

Figure 12. 17 Density of states of L 10 -type MnPt ordered alloy (Sakuma, 2000)
(a) Antiferromagnetic state; (b) Paramagentic ( ~=DLM) state.
566 Akimasa Sakuma, Kazuaki Fukamichi

et al . 2002a). respectively. In each figure. (a) is for the L 10 type tetragonal


structure with c / a = 0.9175. 0.9412 and O. 8796 for MnPt. MnNi and MnPd.
respectively. and (b) in the PM states of corresponding systems. The AFM
magnetic structure of (a) in each figure is coli inear of the AF-I. and the PM
states in (b) are calculated based on the DLM model where the Mn moments
pointing upward and downward are distributed randomly in the Mn sub-lattices
in Fig. 12. 16. It is found that DOS in the AF-I state of each alloy reveals
similar behavior and all have a prominent dip at E F in both spin states. This
characteristic feature can be connected with the AF-I staggered field since the
dip completely vanishes in the PM state as shown in (b) in each figure. These
are closely related to the gap brought about by the AFM long-range order. It
should be noted that the non-magnetic layers such as Pt and Pd layers between
Mn layers interfere opening the full gap and give a small but a finite intensity
at E F
50
40 MnNi(Ll o)
30
~ 20
E
B
OJ
10
.~0- 0 ---=:;;:;;::;:~
;', 10
eo:
~ 20
VJ
o 30
o
40
50 L.-_---'-_ _~_ _- ' -_ _--'-''__----'

-0.8 -0.6 -0.4 -0.2 o 0.2


Energy (Ry)
(a)

40

MnNi(Llo-PM)

g 20
VJ Ni

Mn
40 L-_-:-'--,--_----'-_ _-:-'-:c-_ _: ' - - _ - - , - - J
-0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry)
(b)

Figure 12.18 Density of states of Llo-type MnNi ordered alloy (Sakuma, 1998b)
(a) Antiferromagnetic state; (b) Paramagentic ( ==DLM) state.
Antiferromagnetism of Mn Alloys 567

50
40
30
E
0
20
(;l
to
"
's..
Vl 0
~ 10
-;; 20
0
Ci 30
40
50
-0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry)
(a) MnPd(Ll o)
40

~ 20
(;l

's"..
Vl
0
~
VJ
0 20 Pd
Ci
Mn
40
-0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry)
(b) MnPd(Llo-PM)

Figure 12. 19 Density of states of L 10 -type MnPd ordered alloy (Umetsu et al., 2002a).
(a) Antiferromagnetic state; (b) Paramagentic (~DLM) state.

Besides the shapes of DOS, the magnetic moments of the AF-I and the
PM states are close in each system. In MnPt, for example, the moments are
3. 18 /.Is and 2.94 /.Is in the AF-I and the PM states, respectively, showing that
the reduction of the moment in the PM state is less than 10 %. As far as it
concerns the total energy, the energy in the PM state (b) is larger by as much
as about 88 meV/atom than that in the AF-I state (a) in MnPt, by 119 meV/
atom in MnNi and by 68 meV/atom in MnPd. Figure 12.20 shows the results of
Join the L 10 -type ordered structure for each system (Sakuma, 2001 c). The
value of J o at the actual E F ( = 0) for these alloys is around 150 meV,
comparable to the energy difference between the AF-I and the PM states given
above. The value of TN estimated from J o is about 1000 K, which can be
compared quantitatively with the experimental values. On these grounds, it
seems reasonable to consider that the AF-I state is stabilized through the
formation of a pseudo-gap located at E F' which raises TN at the same time.
568 Akimasa Sakuma, Kazuaki Fukamichi

400

MnPt(Ll o)
200
:>
"
~ 0-- ---
~

-200

-400
-0.6 -0.4 -0.2 0 0.2
EF(Ry)
(a)
400

MnNi(Ll o)
200

:> o-
"
E
~
-200

-400
-0.6 -0.4 -0.2 0 0.2
EF(Ry)
(b)

400

MnPd(Ll o)
200

:>
"E 0
~

-200

-400
-0.6 -0.4 -0.2 0 0.2
EF(Ry)
(c)

Figure 12.20 Effective exchange constant J o of L la-type ordered alloys as a function of


the Fermi level EF(band filling) (Sakuma, 2001c). (a) MnPt; (b) MnNi; (c) MnPd.
Antiferromagnetism of Mn Alloys 569

Let us discuss the magnetic structure in the disordered phase by using the
MnPt system (Sakuma, 2000) for an example. The lattice is fixed as an fcc
structure with the same volume as the ordered phase. The magnetic structures
considered are three multiple-Q states given in Fig. 12. 8 and the PM state
treated by CPA as the DLM state. Figure 12.21 shows DOSs of each magnetic
structure (Sakuma, 2000). It is clear that the pseudo-gap formed in the
ordered phase vanishes and the whole structure changes drastically compared
with the AF-I ordered phase. A highly important aspect is that DOS in the PM
state does not differ so much from these multiple-Q structures. This is a
distinctive feature in the ordered phase, reflected in the total energy difference
as shown in Fig. 12. 22 (Sakuma, 2000). The energy change in the
disordered phase is much smaller than that in the ordered phase. The
difference between the ordered and disordered phases is also much larger than
that of FeMn system and the influence of the entropy is negl igible, which
implies that MnPt system much favors the ordered phase, compared with FeMn
system. We can infer the reason for that from the DOSs as follows. The
random arrangement of Mn and Pt atoms, whose characteristics are much
different from each other, disturbs the AFM staggered field and prevents the

40 40
MnPt (DO-1 Q) MnPt (DO-2Q)
., 20 ., 20
.::
'5,
.::
'5,
'? en
>-. 0 - >-. 0 -
e: e:
:s :s
~ ~

20 20
a a
Mn Mn
40 40
-0.8 -0.6 -0.4 -0.2 0 0.2 -0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry) Energy (Ry)

40 40
MnPt (DO-3Q) MnPt (DO-PM)

<;J 20 ., 20
.::
'5,
.::
'5,
0 -
</> </>
0 -
>-. >-.
e: e:
' 20 ' 20 Pt
a a Mn
Mn
40 40
-0.8 -0.6 -0.4 -0.2 0 0.2 -0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry) Energy (Ry)

Figure 12. 21 Density of states of MnPt disordered alloy for each magnetic structure
(Sakuma, 2000).
570 Akimasa Sakuma. Kazuaki Fukamichi

formation of the pseudo-gap which stabilizes the AFM state. This is in clear
distinction to the case of FeMn system. Even in such a situation. the multiple-
o structures are more stable than the PM state. Among them, the 30 structure
has the lowest energy though the difference from the 20 structure is negligible.
In addition, the 10 structure is higher in energy as much as 5 meV/atom and is
considered to be unstable. In fact, as shown in Fig. 12.23, J o of the 10
structure is slightly negative, which indicates that the 10 structure is not stable
in the ground state (Sakuma, 2000). This may be attributed to the magnetic
frustration of the Mn atoms, which is caused by the random arrangement with
the non-magnetic Pt atoms. On the other hand, the values of J o for the 30 and
20 structures are around 40 meV, being in the region of the energy difference
from the PM state. The value of TN estimated from the value is about 300 K.
The magnetic moment of Mn atoms in the 30 structure is of about 2. 7 J.ls.
smaller than that in the ordered phase as much as about 30 %. In the Mn alloys
with non-magnetic atoms, such as Pt, the results mentioned above lead us to
conclude from that the random distribution remarkably weakens the effects of
staggered field and makes the coli inear magnetic structure unstable because of
the magnetic frustration, which brings about a non-collinear structure close to
the 30 state as well as TN lowered to around room temperature. The situation
is qualitatively the same for MnNi and MnPd alloys in the disordered states.

MnPt
PM

~
IQ 2Q 3Q

Disordered

PM

Ordered

Ll o

Figure 12. 22 Relative difference of the total energy of ordered and disordered MnPt
alloys for each magnetic structure (Sakuma. 2000).
Antiferromagnetism of Mn Alloys 571

150

100 MnPt (DO-I Q)

50
:>
<l.l
E
0
~
-50

-100

-150
-0.6 -0.4 -0.2 0 0.2
EF(Ry)
(a)
150

100 MnPt (DO-2Q)

50
:>
<l.l
0
5
~
-50

-100

-150
-0.6 -0.4 -0.2 0 0.2
EF(Ry)
(b)
150

100 MnPt (DO-3Q)

50
:>
<l.l
0--
5
~
-50

-100

-150
-0.6 -0.4 -0.2 0 0.2
EF(Ry)
(e)

Figure 12. 23 Effective exchange constant J o of MnPt disordered alloys as a function of


the Fermi level E F (band filling) for each magnetic structure (Sakuma, 2000).
572 Akimasa Sakuma, Kazuaki Fukamichi

12.4.4 LIz-Type and y-Phase Mn3Ir Alloys

In the preceding section, we have seen that the L 1a-type Mn ordered alloys
have a high TN' and the random arrangements of the Mn as well as non-
magnetic atoms extremely lower TN' In fact, the deposited Mn alloy films with
1 : 1 composition ratio require heat treatments to realize a sufficiently high
value of the exchange biasing field in GMR and TMR devices. However, such
a heat treatment is unfavorable in industrial processes. In this viewpoint,
alloys having a high TN without any heat treatments are desired for the
complex multi-layered structures in GMR and TMR devices.
Mn-rich system such as y-phase Mnl- x Ir x disordered alloys with x = 0.2-
O. 3 is a promising candidate satisfying the condition mentioned above.
Experimentally, the lattice constants and TN of y-phase Mn,- x Ir x alloys were
first investigated by Yamaoka (1974). He demonstrated that TN of the
disordered alloy at x = 0.25 was around 730 K, very high among Mn-rich alloy
systems. A recent study (Tomeno et aI., 1999) showed that Ll r type Mn31r
ordered alloy also has a high TN of about 950 K. On the other hand, no
systematic theoretical investigations have been performed yet for Mn-rich
alloys, except L 1rtype Mn3 Pt and Mn3 Rh ordered alloys. The magnetic
structure of these two ordered alloys was suggested from both by experimental
(Kouvel and Kasper, 1964; KreJn et al., 1968) and theoretical (KUbler et aI.,
1988) works to be the so-called triangular structure Tl as shown in the inset of
Fig. 12.24. In this sub-section, the electronic and magnetic structures of Mn31r
both of ordered and disordered alloys are presented, and the magnetic
stability will be discussed.
Figure 12.24 shows the local DOS of L l r type Mn31r ordered alloy with
the triangular magnetic structure (Tl) (Sakuma et al., 2003). The lattice
constant is chosen at 3.785 A from the experimental data (Yamaoka, 1974).
The result was drawn from the imaginary part of GF in Eq. (12.25), not from
the usual technique with secular equations, in order to make a comparison with
the later results by CPA for the disordered alloy. It has been confirmed that
this procedure reproduces well both DOS and the moment obtained by the
usual secular equation. The moment of Mn atom is 2.62 fJB and that of Ir is O.
Figure 12.25 shows DOS in a disordered local moment (DLM) state in the PM
state of L l r type Mn31r ordered alloy (Sakuma et al., 2003). The Mn moment
is reduced to 2.20 fJB. whereas the behavior of the local DOS below - O. 1Ry is
retained in both the Tl and the PM states, and the dip around E F found in the
Tl state is swept out in the PM state. This situation is similar to the case of
L la-type MnPt alloy in which the pseudo-gap realized in the AFM state
vanishes completely in the PM state, which means that the gap is closely
Antiferromagnetism of Mn Alloys 573

50
40 Mn]lr (Tl)

30

20
So 10
'2
" 0
'a
Vi
10

~
~
~ 20 ! .
C/]
... jlf
f- . . : ...
g 30 I . I
Mn
40
50 '--_ _...!-_ _--L ' - -_ _---'---_ _----'

-0.8 -0.6 -0.4 -0.2 0.2


Energy (Ry)

Figure 12.24 Density of states of L I, -type Mn3 Ir ordered alloy with a TI state. The
inset shows the triangular structure (TI) of LI, type Mn ordered alloy. The shaded
circles stand for Mn atoms (Sakuma et al , 2003)

connected to the AFM long-range order. The energy in this PM state is higher
than that in the T 1 state by as much as 6 mRy/ atom, giving the same order of
TN' To see the magnetic stability in the T1 state, the effective exchange
constant J o is calculated. Figure 12. 26 shows the value of J o of the Mn
moment as a function of band-filling in the rigid band scheme by scanning E F for
L1z-type Mn31r ordered alloy (Sakuma et al., 2003). It is found that J o has a
maximum value of about 160 meV at the actual E F ( = 0). This result gives TN
of 1250 K, larger by about 30 % than the observed TN' Such a high value of
TN' compared with to the experimental value, would be permissible within the
MFA for the three dimensional system.
40

30 MnJlr (TI-PM)

E 20
0
'2 10
C
.~
0 -
>.
~ 10
~

Vl
0 20
Cl Mn
30

40
-0.8 -0.6 -0.4 -0.2 0.2
Energy (Ry)

Figure 12. 25 Density of states of L 1, -type Mn] Ir ordered alloy in the paramagnetic
(== DLM) state (Sakuma et ai, 2003).
574 Akimasa Sakuma. Kazuaki Fukamichi

300
Mn]lr(TI)
200

100
;;-
<I)
E 0
~
-100

-200

-300
-0.6 -0.4 -0.2 0 0.2
EF(Ry)

Figure 12.26 Effective exchange constant J o of L lz-type Mn31r ordered alloy with a Tl
state as a function of the Fermi level EF'(band filling) (Sakuma et al.. 2003).

Next we proceed to the disordered system. Figure 12. 27 shows DOS of


the 1Q. 2Q and 3Q structures for Mn]lr disordered alloy (Sakuma et al .
2003). The lattice parameter is settled at a = 3 . 785 A. Contrary to the case
of MnPt alloy considered in the preceding Section 12.4.2, the local DOS of Mn
and Ir sites maintains the characteristic features of DOS of the T1 state in the
ordered alloy. Especially, it should be marked that DOS of the 3Q structure is
most close to the T1 state and exhibits a large dip around E F , which is
comparable to the ordered alloy. Therefore, it is considered that the 3Q
structure in the disordered system has a lower energy, compared with the PM
state in the ordered system. Actually, as shown in Fig. 12.28, the 3Q state in
the y-phase disordered alloy is more stable than the PM state in the ordered
alloy (Sakuma, 2001 d). This may be in contrast to the MnPt system where the
MQSDW structures in the disordered alloy have a higher energy than that in the
PM state in the ordered alloy. In the figure, we also plot the energy in the PM
state of the disordered system by using CPA for the random alloy of the
composition Mn( t )0 375 Mn ( t )0 375 lr ( t )0 125 1r ( t )0125. For convenience, the
abscissa is divided into the equal spacing for each magnetic structure. The
energy in the PM state is higher by about 8 mRy/ atom than that in the MQSDW
states, which means that the MQSDW states in the disordered system are
stable, comparable with the T1 state in the ordered system. Figure 12.29
shows the effective exchange constant J o for each MQSDW structure as a
function of the Fermi level E F for Mn]lr disordered alloy (Sakuma et al.,
2003). All exhibit the similar behavior as that in the T1 state in Fig. 12.26,
though the magnitude of the ordinate is sl ightly reduced. Especially, the actual
value of J o at E F = 0 becomes small with a slight shift of the peak position from
Antiferromagnetism of Mn Alloys 575

40
Mn 3Ir(DO-IQ)
E
~ 20
C:
'0..
'">-. -
e:
~
0

(/)
0
a 20 Ir
Mn
40
-0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry)
(a)
40

Mn3Ir(DO-2Q)
E
0
til
20
C:
'0..
'"
~
~
0

(/)
0
a 20 Ir

Mn
40
-0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry)
(b)

40
Mn3Ir(DO-3Q)
E
0 20
til
C:
'0..
'"
~ 0

V;
0
a 20 Ir

Mn
40
-0.8 -0.6 -0.4 -0.2 0 0.2
Energy (Ry)
(c)

Figure 12.27 Density of states of y-phase Mn31r disordered alloys for 1Q, 2Q and 3Q
structures (Sakuma et aI., 2003).

EF=O for the 10 structure. The order of magnitude of J o in these magnetic


structures is consistent with the relative positions of the energy shown in
576 Akimasa Sakuma, Kazuaki Fukamichi

Fig. 12. 28 (Sakuma et al., 2003), and hence the results imply that the 30
structure is most stable in the y-phase disordered system. The value of TN
estimated from J o for Mn31r disordered alloy is about 735 K, in agreement with
the observed value of about 730 K (Yamaoka, 1974), though the present
result for the stable 30 spin structure is different from the 10 spin structure
suggested by Yamaoka (1974). It should be emphasized that powder neutron
diffractions cannot distinguish the 30 structure from the 10 structure as
discussed in the preceding Section 12. 4. 2. The calculated results for both
L 1r type ordered and the y-phase disordered systems are summarized in
Table 12.3 (Sakuma et al., 2003). In addition, J o of the disordered alloy with
a smaller lattice constant of a = 3. 75 A, the same value as y-Mn, does not
change within the numerical accuracy and is still larger than that of y-Mn (Jo =
64 meV). On the contrary, J o of y-Mn with the same lattice constant as that of
the y-phase Mn31r disordered alloy is confirmed to be almost the same as that
of y-Mn with a = 3.75 A. These results disclose that the enhancement of TN by
the addition of Ir to y-Mn is not attributed to the lattice expansion but to the
effect of a change of the electronic structure due to Ir atoms. Note that the
band calculations in the same methods for L 1r type and y-phase Mn3Rh alloys
(Sakuma et a!., 2002) give almost the same results as those for Mn3lr.

PM

PM

Tl

Figure 12. 28 Relative difference in the total energy of L12 -type Mn3 Ir ordered and y-
phase disordered alloys with each magnetic structure (Sakuma et aI., 2003).
Antiferromagnetism of Mn Alloys 577

150
Mn 3Ir(DO-IQ)
100

><l)
50
5 0
..::;
-50

-100

-150
-0.6 -0.4 -0.2 0 0.2
EFCRy)
(a)

150
Mn3[r(DO-2Q)
100

50
><l)

5 0
..::;
-50

-100

-150
-0.6 -0.4 -0.2 0 0.2
Er(Ry)
(b)
150

100 Mn3Ir(DO-3Q)

50
><l)

5 0
..::;
-50

-100

-150
-0.6 -0.4 -0.2 0 0.2
EF(Ry)
(c)

Figure 12. 29 Effective exchange constant J o of y-phase Mn31r disordered alloys as a


function of the Fermi level E F (band filling) for each magnetic structure (Sakuma et aI.,
2003) .
578 Akimasa Sakuma, Kazuaki Fukamichi

Table 12.3 Calculated total energy (I!..E in mRy!atom), the magnetic moment (M in IJs),
the effective exchange constant (J o in meV) and the Neel temperature (na ' in K) estimated
from 2J o!3k s for Mn31r ordered (0) and disordered (DO) alloys (Sakuma et ai., 2003). The
experimental values of the Neel temperature n'p are also shown for comparison.
Phase Spin I!..E MMn M" Jo T~a' T~xp

L lrtype Tl 0 2.62 0.00 162 1253 960(1)


CO) PM 5.7 220 0.00

10 6.9 2.46 0.09 41 317 730(2)


y-phase 20 6.0 2.47 0.11 85 660
(DO) 30 5.6 2.51 0.12 95 735
PM 14.8 2.27 0.00
1999),
(1) (Tomeno et ai.,
(2) (Yamaoka, 1974).

12. 5 Experimental Observations of Antiferromagnetism


of Mn Alloys

As mentioned in the introduction, L 12 -type (-y' -phase) alloys and equiatomic


L l o-type ( CuAu-l-type) Mn alloys have been extensively investigated
because these alloys exhibit a high TN' Therefore, in this section,
fundamental magnetic and electronic properties will be described.

12.5.1 Concentration Dependence of the Neel Temperature of L1 2 -


type (=:..y' -phase) Mn Ordered Alloy Systems
In the y-phase, it has been reported that several kinds of additional elements
such as Cu (Smith and Vance, 1969), Ni (Hicks et aI., 1968), Fe (Endoh and
Ishikawa, 1971) and Pd (Hicks et al., 1968) decrease TN' y-Mnlr, y-MnRh
and y-MnPt alloys easily form an L 1z-type the ordered phase ( y' -phase)
(Raub and Mahler, 1955). In these investigations, Mn-Ir alloy system has
demonstrated that Ir significantly increases TN of y' -phase ordered and y-
phase disordered phases (Yamaoka et aI., 1971, 1974; Yamakoka, 1974).
Therefore, a number of studies for y-phase Mnlr disordered alloys have been
carried out as the exchange biasing films for spin valves.
AFM properties of Mn and its alloys have often been discussed in terms of
the magnitude of the Mn-Mn distance within the framework of the localized
moment model (Yamada et aI., 1970). However, systematic trends of the
concentration dependence of TN are not explained in terms of the model
Antiferromagnetism of Mn Alloys 579

mentioned above. In addition, the magnetic susceptibility of L l r type MnRh


ordered alloys does not follow a Curie-We iss-type temperature dependence as
given in Fig. 12.30, showing no decrease with temperature in PM temperature
ranges (Yamauchi et aI., 1998). In the band theory, in a similar way as that
made in connection with Figs. 12. 15 and 12. 29, the calculations for y-Mn
under the conditions of the collinear spin structure with the lattice constant a =
3.75 A and the axial ratio c / a = 1. 0 provide the exchange energy J o as a
function of the Fermi level E F as shown in Fig. 12.31 (Sakuma, 1998c). As
discussed in the Section 12. 2. 5, the Neel temperature TN is given by the
following expression:

Figure 12. 30 Temperature dependence of magnetic susceptibility for L1z-type MnRh


ordered alloys (Yamauchi et aI., 1998).

(12.41)
The magnitude of J o becomes larger with decreasing the d-electron number,
and hence this figure shows that TN becomes higher with decreasing the d-
electron number at the Mn site. The calculations by the linear muffin-tin orbital
(LMTO)-atomic sphere approximation (ASA) method indicate that the order of
d-electron number at the Mn site is Mn3Rh<y-Mn<Mn3Pt (Sakuma, 1998c).
The concentration dependence of TNshown in Fig. 12.32 (Yamaoka, 1974;
Yamauchi et aI., 1998; Kren et aI., 1968) is consistent with the theoretical
calculations. In other words, the order of magnitude of TN is Mn3Rh > y-Mn>
580 Akimasa Sakuma, Kazuaki Fukamichi

200
y-Mn
1000
100

>'
S" 0 r----~=o:::__--+---+-+---,L-----IO
~

-100 Collinear anti ferromagnetic


a=3.75 A -1000
c/a=1.0
-200
-0.6 -0.4 -0.2 0 0.2
E F (Ry)

Figure 12.31 Exchange constant J o as a function of the Fermi level for y-Mn with a
collinear spin structure (Sakuma, 1998c).

Mn3 Pt (Yamauchi et a!., 1998). In addition, on the basis of Eq. (12.41), it is


expected that not only Ir but also Rh and Ru increase TN' With increasing Rh,
TN increases in a similar manner as Ir, though the value is different in
magnitude. This difference will be discussed in the following Section 12. 6. 1.
L 12 -type ( - y' -phase) MnRu ordered alloys are not obtainable as discussed in
the next Section 12. 5. 2.

1000
~rr
900

800 /Rh
~ 700
~
z
h 600

~
500 y-Mn

400 Pt

300
0 10 20 30 40
x (mol %)

Figure 12.32 Concentration dependence of the Neel temperature TN for L12 -type ordered
alloys; Mnlr (Yamaoka, 1974), MnRh (Yamauchi et aI., 1998) and MnPt (Kren et al.,
1968).
Antiferromagnetism of Mn Alloys 581

12.5.2 Concentration Dependence of the Nee} Temperature of


"{-phase Mn Disordered Alloy Systems

The y-phase Mn-based disordered alloys are stabilized by several kinds of


additional elements in analogy with the y-phase Fe-based alloys. Figure 12.33
shows the concentration dependence of TN of y-phase disordered alloy
systems (Hicks et al., 1968; Endoh and Ishikawa, 1971; Krl'm et al., 1966;
Yamaoka, 1974; Yamauchi et al., 1999; Sasao et al., 1999). The value of
TN exhibits a similar trend with that of the ordered phase given in Fig. 12.32.
What has to be noticed is that the magnitude of TN for the y-phase disordered
alloys is lower by 150 - 200 K than that for L 12 -type ordered counterparts,
depending on the alloying element and its composition. This behavior is
explained by taking randomness into consideration (Sakuma et al., 2002,
2003). Namely, comparing Figs. 12.26 and 12.29 with attention to the fact
that the magnitude of the ordinate is different from each other, the smaller
exchange constant J o is obtained by using the coherent potential approximation
(CPA). It is important to note that TN of y-phase MnRu disorder alloy system
is highest because the d-electron number at the Mn site in MnRu is smaller than
1000

800

g 600
~

y-Mn
400

200 '---~~---'---~~--'~~~--"---~~----'
o 10 20 30 40
x (mol %)

Figure 12. 33 Concentration dependence of the Neel temperature TN for y-phase


disordered alloys; Mnlr (Yamaoka, 1974), MnRh (Yamauchi et al., 1999), MnRu (Sasao
et ai., 1999), MnPt (Kren, 1966), MnPd (Hicks et al., 1968), MnNi (Hicks et ai,
1968) and FeMn (Endoh and Ishikawa, 1971).
582 Akimasa Sakuma, Kazuaki Fukamichi

that in MnRh, in agreement with Eq. (12. 41). Therefore, the smaller the d-
electron number at Mn site, the higher the Neel temperature. However, it is
metallurgically worth noting that the stable y-phase range becomes narrower
with decreasing d-electron number. For example, y-phase MnRu alloys are
unstable and this phase easily transforms into a ~-phase after low temperature
anneal ing (Sasao et aI., 2001). Eventually, in the case of much smaller d-
electron number, say in Mn-Os alloy system, no y-phase is confirmed
(Yamauchi et aI., 2000b; Miyakawa et aI., 2001). From a practical
viewpoint, the stabilization of y-phase MnRu alloys should be accomplished.
For y-phase MnRu alloys, therefore, no excellent exchange-coupled
characteristics are obtained without Rh (Araki et aI., 1998a, 1998b; Tsuchiya
et aI., 2000). By replacing Ru with Rh of about 10%, the y-phase with 0= c
is sufficiently stabilized (Sasao et aI., 2003a). The X-ray diffraction pattern for
Mn70 RU1S Rh s is given in Fig. 12.34. This phase is identified as a y-phase with
0< c (Sasao et aI., 2004). Therefore, y-phase is extended to much lower
concentration ranges by the addition of Rh, compared with Fig. 12. 33, in
which y-phase Mn,oo-. Ru. exists in the range of 31 ~x~37 .

MnsoRulsRhs
a=3.751 A
c=3.813 A

.. ...
50 100 150
2en
Figure 12. 34 X-ray diffraction pattern at room temperature for y-phase Mnao RU15 Rh5
disordered alloy (Sasao et aI., 2004).

12.5.3 Lattice Distortions and Spin Structures of "(-phase Mn


Disordered Alloys
In the preceding Section 12. 4. 1, we have discussed in detail the relation
between the lattice distortion and the spin structure. In this section, we
present experimental concrete examples. Figure 12. 35 shows the phase
diagram of Mn-Ir (Sasao et aI., 2003) disordered alloy system, together with
the data reported by Yamaoka (Yamaoka, 1974). As shown by the dotted
Antiferromagnetism of Mn Alloys 583

800
Para.

600

g 400
h
I
(~) I
200 I
I
a>b>c
,
I

0
Ir(mol %)

Figure 12.35 Phase diagram of y-phase Mnlr disordered alloy system (Sasao et al.,
2003), together with the data by Yamaoka (1974).

line, the Neel temperature TN of y-Mn is extrapolated to be 480 K. On the


whole, the previous composition of the y-phase is sl ightly lower than the
recent data. It is important to note that chemical analyses tend to cause a
slight underestimation of the Ir concentration because the alloys including Ir
atoms are not easy to dissolve comple.tely (Sasao et al., 2003a). The
stabilization of the y-phase below several percent is very difficult,
accompanied by the transformation to a l3-phase. Therefore, the concentration
range of a> b > c in Mn-Ir is strictly restricted to be narrow. For y-phase
Mnlr alloys, the relation between the electronic total energy and the spin
structure has been investigated (Sakuma et al., 2003). The spin structure
changes from the 20 to 30 state around 13 % Ir under the condition of c/ a = 1.
Emphasis to be made in Fig. 12. 35 is that the Ir concentration where the
magnetic transition from the 20 to 30 structure takes place is quite close to the
concentration at which the axial ratio changes from c/ a > 1 to c/ a = 1. It has
been predicted that even in the pure y-Mn metal, the lattice distortion of
c / a > 1 further lowers the total energy of the 20 structure (F ig. 12. 10). In a
similar manner as above, the total energy of the 20 structure in y-phase
MnB51r15 disordered alloy becomes lower than that of the 30 structure when the
lattice is distorted to c/ a = 1.05, whereas the 30 structure in y-phase MnBO
=
Ir20 disordered alloy is still stable under the condition of c / a 1. Thus, it can
be inferred that the critical concentration defined with X m for the magnetic
transition is put forward to a higher concentration in a few percent if the
constraint for the lattice distortion is free. It should be marked, however, that
the concentration X m does not necessarily coincide with the concentration
defined by x t for the lattice distortion. This can be understood by noting a
situation that y-phase MnBO Ir20 disordered alloy having the 30 structure is found
to prefer c/ a > 1 energetically rather than c/ a = 1, while for y-phase
584 Akimasa Sakuma, Kazuaki Fukamichi

Mn7sIr2s disordered alloy the lattice keeps cia = 1 with the 30 structure. From
these relations, one can expect that x t is higher than x m' This is because the
lattice distortion of cia> 1 lowers not only the total energy of the 20
structure, but also that of the 30 structure. When the total energy of the 30
structure is low enough in c I a = 1, any degree of the lattice distortion cannot
always cause a reversal from the 30 to 20 magnetic structure (Sakuma et al.,
2003). Fishman and Liu (1999) have calculated theoretically by using the
random phase approximation, the temperature dependence of magnetic
susceptibility for three kinds of the spin configurations illustrated in Fig. 12.8,
namely, the single (10), double (20), and triple (30)-spin density wave
(SOW) structures. Following their theory, we can determine the spin structure
change. The temperature dependence of magnetic susceptibility of the
y-Mnl00- x Ir x disordered alloys with x = 15. 5 and 14. 8 is shown in Figs. 12.
36a and 12. 36b (Sasao et aI., 2003). At the transformation temperature from
the high-temperature 30 spin structure to the low-temperature 20 spin
structure, T 2Q / 3Q , the susceptibility suddenly increases or decreases by a few
percent, depending on the size difference between the electron and whole
Fermi surfaces (Fishman and Liu, 1999). It should be noted that a clear
decrease is observed below T 2Q/3Q as seen from Fig. 12. 36a, whereas no
anomaly in the temperature dependence of magnetic susceptibility is observed
at the crystallographic transition temperature from the fcc to the fct phase,

7.5
y- Mn 852 1r t48
~ 7.0
OJ)

-=E 6.5
","
b
x 6.0
~

~
5.5
0 200 400 600 800
T(K)
(a)
7.5
Y- Mn 845 lr I55
Oil 7.0 TN
-=E
"," 6.5
b
~ 6.0
~

5.5
0 200 400 600 800
T(K)
(b)

Figure 12.36 Temperature dependence of the magnetic susceptibility of (a) y-Mn8521r148


and (b) y-Mn8ulrt55 disordered alloys (Sasao et ai., 2003).
Antiferromagnetism of Mn Alloys 585

TIel/Ice' as given in Fig. 12. 36a, b. Accordingly, we ascertain that the lattice

distortion is not accompanied by the spin structure change at the same time.
The phase diagram of Mn-Rh alloy system in Fig. 12.37 (Yamauchi et aI.,
2000a) is similar to that of Mn-Ir alloy system given in Fig. 12.35, but an
additional phase with a> c appears in Mn-Rh system. The relation between
the lattice distortion and the spin structure of y-phase MnRh alloys is shown in
Fig. 12.38 (Yamauchi et aI., 2000b). In the case of a = c, the spin structure
is 30, and the other structures depend on the lattice distortion. It should be
emphasized that the spin structure of MngoRh lO with a>b>c in Fig. 12.38 is
consistent with the neutron diffraction data (Hori et al., 2001) and also
theoretical calculations (Jo and Hirai, 1986).
800
Para. a=c

600

a=c
200

o 10 20 30
Rh (mol %)

Figure 12.37 Phase diagram of y-phase MnRh disordered alloy system (Yamauchi et aI.,
1999).

SDW IQ 2Q 3Q

Lattice a>c a<c a>b>c a=c

Moment /1?- /1a = /1h =0 /1a=/1h>/1c=O /1?- /1b>/1a=0 /1a=/1h=/1c

Structure
[] [J []
J------ []
Figure 12.38 Lattice distortion, the magnetic moment and the spin configuration in 10,
20 and 30 states for y-phase MnRh disordered alloys (Yamauchi et aI., 2000a).

It has been pointed out that excellent spin valve characteristics of y-phase
Mnlr alloys are not obtainable below around 20 % Ir (Hoshino et aI., 1996;
Fuke and Kamiguchi, 1998), very close to the composition in which the lattice
586 Akimasa Sakuma, Kazuaki Fukamichi

distortion takes place as seen from Fig. 12. 35, (Sasao et al., 2003b),
suggesting that exchange biasing characteristics are sensitive to the spin
structure. As discussed in Sections 12. 4. 2, 12. 4. 3 and 12. 4. 4, the spin
structure of MnPt resembles that of FeMn with increasing Mn content, or
coming close to the y-phase. This strongly implies that the spin structure of
y-Mn disordered alloys is easily modified by lattice defects, magnetovolume
effects and so on, because the energy difference between 1Q, 2Q and 3Q
states is not so significant (Figs. 12.13 and 12.28).

12.5.4 Concentration Dependence of the Nee) Temperature of


L1 0 ( CuAu-I)-Type Alloy Systems

In the vicinity of equiatomic composition, TN of several kinds of L1 rtype alloy


systems is high enough to use for GMR and TMR devices. Illustrated in
Fig. 12.39 is the concentration dependence of TN of MnNi (Kjekshus and
M6l1erud, 1967), MnPt (Krem et ai., 1968) and MnPd (Pal et ai., 1968),
together with that of Mn-Cr (Hamaguchi and Kunitomi, 1964) and Fe-Mn
(Endoh and Ishikawa, 1971). In the former three kinds of alloy systems, a
high temperature phase has a CsCI-type structure and a martensitic phase
transformation induces a low temperature phase with L 10 (= CuAu-1 )-type
structure. L 10 -type MnPt alloys are widely used as exchange biasing films of
1200

1000

800

g
K 600

400

200 '------''-----'_--'-_--'-_-----L_---'-_----'
W ~ ~ w w W W 100
x (mol %)

Figure 12. 39 Concentration dependence of the Neel temperature TN for L10 -type Mn
alloy systems; MnPd (Pal et al., 1968; Kjekshus and Mollerud, 1967>, MnNi (Kren
et aI., 1968) and MnPt (Kren et aI., 1968; Pal et aI., 1968), together with that of FeMn
(Endoh and Ishikawa, 1971) and CrMn (Hamaguchi and Kunitomi, 1964) alloy systems.
Antiferromagnetism of Mn Alloys 587

spin valves. Generally speaking, MnPt sputtered films are in a disordered


state and non-magnetic at room temperature, and hence the heat-treatment for
ordering is necessary to obtain the AFM state. Compared with Fe-Mn alloy
system, the concentration dependence of TN of L1 a-type MnPt and MnPd alloy
systems is significant, consistent with the theoretical calculations given in
Section 12.4. 3. Figure 12. 40 shows the Neel temperature TN versus the d-
electron number in L 1a-type Mn alloys, together with that of L 1rtype ordered
and y -phase disordered Mn alloys (Sakuma, 1999; Fukamichi, 2000). The
solid and open squares stand for the experimental and calculated results,
respectively. On the whole, TN decreases with an increase of the number of d-
electrons. Recently, Wang et al. (2001) have carried out the first principles
calculations on MnAu superlattices and pointed out that TN increases with
decreasing the lattice tetragonality cia. The magnitude of TN of Ll a-type
alloys given in Fig. 12.40 is consistent with their calculations because c I a is
0.94 for MnNi, 0.92 for MnPt and 0.88 for MnPd.
o
y- Mn 3 Pt

-500 y-Mn (,if)


." -
(disordered) - y- Mn3 Pt
(ordered)
..... y- Mn 3Rh
-"-., (disordered)
----- (ordered)
g -1000 (~'::
..... MnPt
~
I

-1500

-2000 L-_----'-_ _- ' -_ _' - - _ - - - ' -_ _---'


5.2 5.3 5.4 5.5 5.6 5.7
Number of d-electrons

Figure 12.40 The Neel temperature TN as a function of number of d-electrons for L10 -type
Mn alloys, together with that of y-Mn and y-phase Mn disordered alloys (Sakuma, 1999;
Fukamichi, 2000). The solid and open squares, and 0, stand for the experimental and
calculated results, respectively.

12.5.5 Low-Temperature Specific Heat and Temperature


Dependence of Electrical Resistivity

The electronic states in L la-type and y-phase Mn alloy systems have been
discussed in the preceding Section 12.4. It is characteristic of L la-type MnPt,
588 Akimasa Sakuma, Kazuaki Fukamichi

MnPd and MnNi alloys to have a pseudo-gap in the electronic states. The low-
temperature specific heat and the temperature dependence of electrical
resistivity provide useful information on the electronic states. Figure 12. 41
shows the low-temperature specific heat C in the conventional form of C / T
vs. T 2 for L l r type MnRh ordered alloys (Umetsu et aI., 2003) and L l o-type
MnPd, MnPt and MnNi alloys (Umetsu et al., 2002a, 2002b). The specific
heat coefficient proportional to the DOS is obtained by a linear extrapolation to
T = O. The value of the latter L 10 -type alloys is extremely small, compared
with that of the former L 1rtype ordered alloys, consistent with the band
calculations given in Figs. 12.17, 12.18 and 12.19.
15

~ 10
"0
y
E -0-
MngoRh zo
-,
5
h
U
5
---
~

--cr-
Mn 3 Rh
MnPd
MnPt
-0- MnNi

o 10 40 50

Figure 12.41 Low-temperature specific heat C as a function of temperature T in the form


of C / T vs. T 2 for L1 rtype MnRh ordered alloys (Umetsu et aI., 2003), together with
that for L 10 -type Mn alloys with Ni, Pd and Pt (Umetsu et aI., 2002a; Umetsu et aI.,
2002b) .

The temperature dependence of the electrical resistivity of L 12 ( -


y' -phase) MnRh ordered alloys (Umetsu et aI., 2003) and L 10 -type MnPt
alloys (Umetsu et aI., 2002b) is presented in Figs. 12. 42 and 12. 43,
respectively. The former exhibits a monotonic decrease below TN' which is
associated with a usual magnetic order-disorder scattering, whereas the latter
shows a pronounced hump below TN' being characteristic of band gap-type
antiferromagnets. The other L l o-type alloys also exhibit similar temperature
dependence (Umetsu et al., 2002b). Accordingly, the experimental data in
Figs. 12.42 and 12.43 for L 10 -type Mn alloys are in support of the theoretical
calculations.
Antiferromagnetism of Mn Alloys 589

250

200

E
'?
:, 150 MnS3Rh l7
--' J
- - .........

Q
......\, ~ .
.

100

50 L -_ _---'- --'----_ _-----.JL-_ _---'

300 500 700 900 1100


T(K)
Figure 12. 42 Temperature dependence of the electrical resistivity of L1rtype MnRh
ordered alloys <Umetsu et aI., 2003).

150

~100
E
u

~
Q 50
MnPt
(Llo-type)

o 200 400 600 800 1000 1200


T(K)
Figure 12. 43 Temperature dependence of the electrical resistivity for L l o-type MnPt
alloys <Umetsu et aI., 2002b).

12. 6 Antiferromagnetic Properties Related to


Magnetic Devices

For advanced electronic and magnetic devices, such as GMR and TMR
devices, various kinds of AFM properties should be controlled. Detailed data
on these data will be discussed by J. X. Shen et ai., and H. H. Garzen and
P. Grunberger in the separate series of Advanced Magnetic Materials.
Therefore, in this section, we restrict ourselves within the discussions
concerned with the Neel temperature and the spin structure.
590 Akimasa Sakuma, Kazuaki Fukamichi

12. 6. 1 Magnetovolume Effects and Thermal Strains


Several kinds of the thermal expansion curves of L 1z-type ordered alloys are
drawn in Fig. 12.44 (Kren et aI., 1968; Tomeno et aI., 1999; Umetsu et aI.,
2003). These curves exhibit a clear shrinkage below TN' indicating a negative
spontaneous volume magnetostriction ws. The thermodynamic Eherenfest
equation is given by
(12.42)

200 400 600 800 1000


T(K)
Figure 12.44 Thermal expansion curves of L12-type ordered alloys; Mn82 Pt 18 (Krim et
ai., 1968), Mn80 Rh20 and Mn3 Rh (Umetsu et ai., 2003) and Mn31r (Tomeno et ai., 1999).

where V and P are the volume and pressure, respectively, /::i.av and /::i.C m are
the difference between the thermal expansion coefficient and the specific heat
before and after of the magnetic phase transition, respectively. The equation
indicates that the negative Ws increases TN with applying hydrostatic
pressure. Figure 12.45 shows the pressure effect on TN for L1z-type MnRh
ordered alloys. With increasing pressure, or decreasing the lattice constant,
TN increases in accordance with Eq. (12.42). The concentration dependence
of the lattice constant of y-phase is given in Fig. 12. 46 (Sasao et aI., 1999;
Antiferromagnetism of Mn Alloys 591

Yamaoka, 1974; Yamauchi et a!., 1999; Krln, 1966; Kren et a!., 1968).
Noteworthy is that the value of MnRh is larger than that of Mnlr. In Fig. 12.47,
TN of Mnlr is higher than that of MnRh, though Ir and Rh have the same valence
electrons. This difference would be explained from the data in Figs. 12.45 and
12.46. That is to say, the shrinkage of the lattice constant brings about the
increase of TN' From the pressure effect on the lattice constant for L 12 -type
MnRh ordered alloys in Fig. 12. 45, the compressibility of L 1r type (=
y' -phase) Mn3 Rh ordered alloy is estimated to be 1. 4 x 1O- 2 /GPa, much
larger than that of Ta , Mo , Nb and V . The temperature dependence of the

900
dTN/dP=7.0 K/GPa
880

860

Q
~ 840 dTN/dP=8.8 K/GPa
~

820

2 3 4 5 6
P(GPa)

Figure 12. 45 The Neel temperature TN as a function of pressure for L1 2 -type MnRh
ordered alloys (Umetsu et ai., 2003).

3.86

3.84

0$
-

Vl
3.82

c
8 3.80
8
.~
.....J
3.78
Mn-Ru
...... ..-e

3.76 ' - - - - - - - - - ' - - - - - - ' - - - - - - '


10 20 30 40
x (mol%)

Figure 12. 46 Concentration dependence of the room temperature lattice constant for
y-phase disordered alloys; MnRu (Sasao et al., 1999), Mnlr (Yamaoka, 1974), MnRh
(Yamauchi et ai., 1999) and MnPt (Kren et ai., 1968).
592 Akimasa Sakuma, Kazuaki Fukamichi

lattice constant a for L1 o-type MnPt alloy is given in Fig. 12.47 (Pal et aI.,
1968). A remarkable shrinkage also occurs below TN' Therefore, it is
expected that the shift of TN against pressure for L1 0 -type Mn alloys is in
analogy with that of L 1r type Mn ordered alloys.
2.5

2.0 MnPt

~
1.5
~
~
<:l
<I 1.0

0.5

0 1500
T(K)
Figure 12.47 Temperature dependence of the lattice constant for Ll o-type MnPt alloy
(Pal et ai, 1968).

100

80

:,.:,..;:.

\~.:.:'::',~
,I .................; . .......

20

o 200 400 600 800 1000


T(K)
Figure 12.48 Thermal expansion coefficient as a function of temperature for Ll r type
Mn3 Rh ordered alloy (Umetsu et ai, 2003).

The temperature dependence of the thermal expansion coefficient of L 1r


type Mn3Rh ordered alloy is presented in Fig. 12.48 (Umetsu et aI., 2003).
The thermal expansion coefficient ex is given by
(12.43)
where ex ph' ex el and ex m are the phonon, the electron, and the magnetic terms,
Antiferromagnetism of Mn Alloys 593

respectively. The solid line in the figure is aph + a.I' which is estimated from
the Debye model by using low-temperature specific heat data. Note that the
thermal expansion coefficient in the PM temperature ranges is very large of
about 30 x 10- 6 K- 1 , compared with conventional 3d metals and their alloys,
deviating from the solid curve. The enhancement of the thermal expansion
above the magnetic transition temperature is explained by the spin fluctuation
(SF) theory (Moriya, 1985). According to this theory, the spontaneous
volume magnetostriction Ws at T is expressed by
(12.44)
where K, C, M, and 52 ( T) are the compressibility, the magnetovolume
coupl ing constant, the magnetization, and the mean square ampl itude of
thermal SFs, respectively. This figure presents that L 1r type Mn3 Rh ordered
alloy exhibits a marked SF effect over a wide temperature range even above
TN'
Spin valve devices are constructed with mutilayers. In making processes
of the multilayers, the substrate temperature would become high above room
temperature. In such circumstances, it is expected that thermal strains are
inevitably induced at a high level in the mutilayers due to the difference of the
thermal expansion coefficient among the film layers. From practical
viewpoints, Ws and thermal strains are not ignored because they are closely
correlated with durability and stability of spin valves. Namely, these strains
easily accelerate diffusions associated with electromigration and
stressmigration in spin valves, causing deterioration of spin valve
characteristics.

12. 6. 2 Blocking Temperature and Magnetic Domains in


Antiferromanets
In the case of ferromagnets, magnetic domains are easily formed in order to
reduce the magnetostatic energy. On the contrary, generally, antiferrmagnets
have no magnetostatic energy as expected from a simple AFM spin structure,
and hence no magnetic domains would exist in ideal antiferromagnets.
However, there are many defects and strains in real crystals, and then the
domains with finite sizes are formed in antiferromagnets. Usually, the
observation of AFM domains is not so easy by conventional methods such as
optical, X-ray and neutron techniques because the observations have been
carried out through a very small lattice distortion caused by the exchange
strain below the magnetic transition temperature. It has been considered that
the AFM domain size is smaller than the crystal grain size. In line with such an
assumption, many published papers have discussed the relation between the
grain size and the strength of exchange coupling and also the blocking
temperature (T B ) . At the present stage, however, the discussions are
594 Akimasa Sakuma, Kazuaki Fukamichi

divergent on this point.


Recently, a powerful method for the observation of the AFM domains has
been developed. By using X-ray magnetic linear dichroism (XMLD)
spectroscopy, it has been demonstrated that the domain size in NiO film grown
on Mg( 100) is smaller by 1 or 2 orders of magnitude than that in a bulk NiO
(Stohr et al., 1999). On the other hand, the correspondence between the
blocking temperature T B and the Neel temperature TN has often been
discussed by many researchers. Associated with these discussions, the AFM
domain observation of LaFe03 is meaningful. It should be pointed out that the
domains observed by XMLD spectroscopy for LaFe03 grown epitaxially on
SrTi03(100) thoroughly disappears at a lower temperature by about 70 K than
TN of the bulk specimen. Thermal strains bring about this behavior by changing
the angle of the Fe-O-Fe superexchange interaction (Scholl et ai., 2000), and
hence reduce TN' This data clearly show that the magnetic transition
temperature is sensitively affected by thermal strains. Direct observations of
the domain structures in Mn alloy systems are very important to develop
excellent GMR and TMR devices because the thermal strains and the
magnetovolume effects presented in Section 12.6.1 are liable to change TN as
well as the spin structure discussed in Section 12.5.3.

12.6.3 Frustration of Antiferromagnetic Spin Structures and


Exchange Bias Field
In contrast to ferromagnetic (FM) materials, the frustration of spin structures
in AFM materials plays an apparent role in magnetic properties. In Fig. 12.8,
the frustration makes the 1Q structure unstable, and therefore another non-
collinear structure is stabilized, resulting in a lower TN' Macroscopically, the
frustration would also be induced by crystallographic defects as well as off-
stoichiometry. In this section, the AFM spin structures at the interface
between AFM and FM bilayers are discussed.
Many materials to gain the exchange bias field have been developed, and
also many models have been proposed to explain the mechanism of the
exchange bias. In these models, two major problems are pointed out; first,
the quantitative disagreement of the magnetic energy which is much smaller
than the interatomic exchange interaction. Second, the cancellation of
spins S at the interface of the AFM layer, that is, the so called
compensated interface. Introducing the magnetic domain wall structure near
the interface, the quantitative agreement was first achieved by Mauri et al.
( 1987). Both the quantitative agreement and the problem of the compensated
interface were explained within the framework of a random field model by
Malozemoff (1987). In this model, both the roughness at the interface and the
Antiferromagnetism of Mn Alloys 595

AFM domain are required. On the other hand, Koon (1997) proposed a spin
flop model having a smooth interface, and explained relaxed spin structures at
the interface. Later on, however, the different solutions in the spin flop
arrangement for the magnetization reversal process were pointed out
(Schulthess and Butler, 1998; Xi and White, 2000). Those models are based
on the collinear spin structure which shows the antiparallel arrangement of AFM
spins in the AFM layer. However, it should be pointed out that various types of
Mn alloys are used in industrial applications, and their spin structures are not
necessarily the collinear spin structure as discussed in Section 12.4.
Assuming a bilayer constructed by the < 111> stacking in the triangular net
plane, forming a compensated interface between the FM and AFM layers, the
Monte Carlo calculations within the framework of the classical Heisenberg
model have been carried out (Mitsumata and Sakuma, 2001 ; Mitsumata et aI.,
2003a, 2003b). For the magnetic structures in the AFM layer, the MSDW for
y-phase disordered alloys and AF-I structure for L l o-type ordered alloys were
considered by taking appropriate contents of non-magnetic atoms in the fcc
lattice. In the classical Heisenberg model, it should be noted that the 30
structure is essentially preferred among the MSDW structures due to the
existence of randomly arranged non-magnetic atoms in the AFM fcc lattice.
The 10 and 20 and AF-I structures can be realized by making the exchange
interactions anisotropic; different values between in the inter-c-plane and the
intra-c-plane. The calculated magnetization loops are presented in Fig. 12. 49.
We observe that only the 30 structure can create the exchange coupling bias
instead of the 1, 20 and AF-I structures. These results tell us that the non-
collinear spin structure caused by the geometric frustration in the AFM layer is
responsible for the magnetization loop shift. It is meaningful to note that the
exchange bias field is drastically reduced with decreasing Ir concentration
below about 20%lr (Hoshino et aI., 1996; Fuke and Kamiguchi, 1998), very
close to the concentration where the spin structure change from the 30 to 20
spin structure takes paces (Fig. 12. 35). In addition, epitaxial growth
characteristics of Mnlr films drastically change around 20 % Ir (Fuke et al.,
1997) On the other hand, a collinear spin structure formed in AFM L l o-type
ordered alloys results in only the enhancement of coercivity of the FM layer.
IntrOducing the multi-domains into the ordered L 10 -type AFM layer, however,
the magnetization loop shift of the FM layer is evidently developed by the
geometrically frustrated spins induced at the magnetic domain boundaries
(Mitsumata et al. 2003a). It is important to note that the magnitude of the
exchange bias field corresponds to the gain of the exchange energy due to the
canting of spins caused by the frustration. This explains that L l o-type ordered
alloys having a large canting angle of spins exhibit a large exchange bias field
(Lin et aI., 1994; Saito et aI., 1997), compared with that of the y-phase
disordered alloys. In this frustration model, the introduction of the frustrated
596 Akimasa Sakuma, Kazuaki Fukamichi

spins configuration in the AFM layer and the formation of the magnetic domain
wall parallel to the interface of the FM and AFM bilayer are necessary for the
development of the unidirectional exchange bias field. These prerequisites,
the frustrated spins and the domain wall, are common to both the disordered
y-phase and ordered L 10 -type AFM systems. In addition, the multi-domain
structures to create the frustrated spins are also required for the collinear spin
structure in the ordered L1 0 -type AFM layer. The AFM domains structures are
in part similar to the random field model (Malozemoff, 1987). The random
field model stands within the framework of the Ising spin system. In a wide
sense, however, the AFM domain in the random field model can be regarded
as the frustration model. The domains also correspond to the grains in the
polycrystalline models (Nishioka et ai., 1996; Tsunoda et ai., 1997) Those
two kinds of the polycrystalline models include no influences of the frustrated
spins at the interface. As a consequence, those polycrystalline models are
naturally extended to the present model of the compensated interface by taking
the influence of frustrated spins into consideration.
1.5
I

1.0 r
: I
I I
I
~ I I I
Vl 2Q ..... I I
~ 0.5
I

~
Ii..........
I
I
I
I
I
I ..........AF-I
I
c: I I I
I
0
0 ----- ...,-- I
-+ ----~I ------
:~v I
I
I
I
I
I
I
I
I
c: I
bJj -0.5
oj
3Q-
-t---- I ~
~ I I
I
I
I : ' " IQ
-1.0 _J
I

-1.5 :
-6 -3 0 3 6
Applied field (gJ.1 B H/D)

Figure 12.49 Magnetization loops of AFM/FM bilayer with various spin structures in AFM
layer CMitsumata et aI., 2003b).

As discussed above, the role of the spin frustration caused by the


competition of exchange interaction due to the geometry of the magnetic atoms
in the AFM disordered layer is significant. From the frustration model, it
possibly explains the mechanism for the dependence of the exchange coupling
bias on the composition in AFM fcc alloys (Mitsumata et ai., 2003b). If the
origin of the exchange coupling bias is the interface roughness, the interface
roughness should change with the composition in the AFM layer, or else the
dependence of the exchange coupling bias on the composition in the AFM layer
is hardly explained.
Antiferromagnetism of Mn Alloys 597

12.7 Concluding Remarks

For the FeMn alloy system, both the electronic state and its energy are not so
sensitive to the atomic arrangement of Fe and Mn, whereas the spin
configuration is compl icated. Especially the energy difference between 1Q,
2Q and 3Q in the disordered state is very small. On the other hand, the
stability of the AFM structure in L la-type alloys is closely correlated with the
ordered atomic structure and the AFM state is stabilized when the staggered
field caused by the Mn atomic plane makes a pseudo-gap at the Fermi level.
In contrast to L 1a-type alloys, the frustration among the Mn moments in
y-phase disordered and L l r type ordered alloys plays an important role ,
associated with the non-collinear spin structure. The experimental data on the
magnitude of TN is explained by taking the electron concentration of the Mn
site into consideration.
The energy difference between the spin structures in the y-phase Mn
alloys is not so large, and hence stresses as well as strains would influence on
the AFM properties. The magnetovolume effects and also crystallographic
defects result in stresses and strains. These facts allow us to speculate that the
spin structure in AFM films of spin valves is locally different, depending on the
magnitude of induced strains. Therefore, practically, direct observations of
AFM domains in Mn-alloys are very important in order to accomplish excellent
spin valve characteristics.
Finally, we should stress that magnetic and other fundamental data for
antiferromagnets amount to little, compared with those for practical
ferromagnets. For further progresses of advanced magnetic devices,
accumulation of these data is earnestly desired.

References
Akbar, S., Y. Kakehashi and N. Kimura. J. Phys. Condens. Matter 10: 2081
(1998)
Andersen, O. K. Phys. Rev. B 12: 3060(1975)
Antropov, V. P., M. I. Katsunelson, B. N. Harmon, M. van Schilfgaarde and
D. Kusnezov. Phys. Rev. B 54: 10 19( 1996)
Araki, S., E. Omata, M. Sano, M. Ohta, K. Noguchi, H. Morita and M.
Matsuzaki. IEEE Trans. Magn. 34: 387 ( 1998a)
Araki, S., M. Sano, M. Ohta, Y. Tsuchiya, K. Noguchi, H. Morita and M.
Matsuzaki. IEEE Trans. Magn. 34: 1426( 1998b)
598 Akimasa Sakuma, Kazuaki Fukamichi

Asano, S. and J. Yamashita. J. Phys. Soc. Jpn. 31: 1000( 1971)


Bisanti, P., G. Mazzone and F. Sacchetti. J. Phys. F17: 1425(1987)
Cade, N. A. and W. Young. J. Phys. F: Met. Phys. 10: 2035(1980)
Endoh, Y. and Y. Ishikawa. J. Phys. Soc. Jpn. 30: 1614(1971)
Fishman, R. S. and S. H. Liu. Phys. Rev. B 59: 8672(1999)
Fujii, S., S. Ishida and S. Asano. J. Phys. Soc. Jpn. 60: 4300(1991)
Fukamichi, K. Materia Japan (in Japanese) 39: 790(2000)
Fuke, H. N., K. Saito, Y. Kamiguchi, H. Iwasaki and M. Sahashi. J. Appl.
Phys. 81: 4004( 1997)
Fuke, H. N. and Y. Kamiguchi. J. Magn. Soc. Jpn. 22: 58(1998)
Hamaguchi, Y. and N. Kunitomi. J. Phys. Soc. Jpn. 19: 1849 ( 1964)
Hicks, T. J., A. R. PepperandJ. H. Smith. J.Phys. Cl: 1683(1968)
Hirai, K. and T. Jo. J. Phys. Soc. Jpn. 54: 3567(1985)
Hori, T., S. Tanaka, Y. Tsuchiya and K. Houjo. J. Phys. Soc. Jpn. 70:
Suppl. A 142(2001)
Hoshino, K., R. Nakatani, H. Hoshiya and Y. Sugita. Jpn. Appl. Phys. 35:
607(1996)
Hubbard, J. Phys. Rev. lett. 3: 77( 1959)
Hubbard, J. Proc. Roy. Soc. london. A 276: 6238 (1963)
Hubbard, J. proc. Roy. Soc. london. A 277: 237 (1964)
Hubbard, J. proc. Roy. Soc. london. A 281: 401 (1964)
Janak, J. F. SolidStateCommun. 25: 53(1978)
Jo, T. and K. Hirai, J. Phys. Soc. Jpn. 55: 2017(1986)
Kasper, J. S. and J. S. Kouvel. J. Phys. Chem. Solids 11: 231(1959)
Kennedy, J. and T. J. Hicks. J. Phys. F 17: 1599( 1987)
Kjekshus, A., R. Mollerud., A. F. Anderson, and W. B. Pearson. Phil.
Mag. 16: 1063( 1967)
Korenman, V., J. l. Murrag and R. E. Prange. Phys. Rev. B 16: 4032
( 1977)
Koon, N. C. Phys. Rev. lett. 78: 4865(1997)
Kouvel, J. S. and J. S. Kasper. Proc. Int. Conf. on Magn. Nottingham, P.
169 (1964)
Kren, E., G. Kadar, l. Pal, J. S6lyom, and P. Szab6. Phys. lett. 21: 383
(1966)
Kren, E., G. Kadar, l. Pal, J. S6lyom, P. Szab6 and T. Tarn6czi. Phys.
Rev. 171: 574( 1968)
KObler, J., K.-H. Hock, J. Sticht and A. R. Williams. J. Phys. F: Met.
Phys. 18: 469( 1988)
Kudrnovsky, J. and V. Drchal. Phys. Rev. B 41: 7515 (1990)
Liechtenstein, A. I., M. I. Katsnelson and V. A. Gubanov. Solid State
Commun. 54: 327 ( 1985)
Liechtenstein, A. I., M. I. Katsnelson, V. P. Antropov and V. A. Gubanov.
Antiferromagnetism of Mn Alloys 599

J. Magn. Magn. Mater. 67: 65(1987)


Lin, T., D. Mauri, N. Staud, C. Hwang, J. K. Howard and G. Gorman.
Appl. Phys. Lett. 65: 1183( 1994)
Malozemoff, A. P. Phys. Rev. 835: 3679 ( 1987)
Mauri, D., H. C. Siegmann, P. S. 8agus and E. Kay. J. Appl. Phys. 62:
3047( 1987)
Meneghetti, D. and S. S. Sidhu. Phys. Rev. 105: 130 ( 1957)
Mitsumata, C. and A. Sakuma. J. Magn. Soc. Jpn. (in Japanese) 25: 823
(2001 )
Mitsumata, C., A. Sakuma and K. Fukamichi. Phys. Rev. 8. phys. Rev. 8
68: 014,437(2003a)
Mitsumata, C., A. Sakuma and K. Fukamichi. IEEE Trans. Mag. 39: 2738
(2003b)
Miyakawa, M., R. Y. Umetsu and K. Fukamichi. J. Phys. Condens. Matter.
13: 3809(2001)
Moriya, T. Spin Fluctuations in Itinerant Electron Magnetism. Springer-
Verlarg( 1985)
Nishioka, K., C. Hou, H. Fuj iwara and R. D. Metzger. J. Appl. Phys. 80:
4528 (1996)
Pal, L., E. Kren, G. Kadar, P. Szabo and T. Tarnoczi. J. Appl. Phys. 39:
538(1968)
Raub, V. E. and W. Mahler. Z. Metallkde. 46: 282(1955)
Saito, M., Y. Kakihara, T. Watanabe and N. Hasegawa. J. Magn. Soc.
Jpn. 21: 505(1997)
Sakuma, A. J. Phys. Soc. Jpn. 67: 2815(1998a)
Sakuma, A. J. Magn. Magn. Mater. 187: 105 (1998b)
Sakuma, A. unpublished (1998c)
Sakuam, A. IEEE Trans. Mag. 35: 3349(1999)
Sakuma, A. J. Phys. Soc. Jpn. 69: 3072 (2000)
Sakuma, A., unpublished (2001a)
Sakuma, A., unpublished (2001b)
Sakuma, A., unpublished (2001 c)
Sakuma, A., unpublished (2001 d)
Sakuma, A. R. Y. Umetsu and K. Fukamichi. Phys. Rev. B 66: 014,432
(2002)
Sakuma, A. K. Fukamichi, K. Sasao and R. Y. Umetsu. Phys. Rev. 8 67 :
024,420 (2003)
Sasao, K., R. Yamauchi, K. Fukamichi and H. Yamauchi. IEEE Trans.
Magn. 35: 3910 (1999)
Sasao, K., R. Y. Umetsu and K. Fukamichi. J. Alloys Compds. 325: 24
(2001 )
Sasao, K., R. Y. Umetsu and K. Fukamichi. J. Alloys Compds. 352: 21
600 Akimasa Sakuma, Kazuaki Fukamichi

(2003)
Sasao, K., R. Y. Umetsu , K. Fukamichi and A. Sakuma. J. Alloys
Compds. in press(2004)
Scholl, A., J. Stohr, J. Luning, J. W. Seo, J. Fompeyrine, H. Siegwart,
J. -P. Locquet, F. Nolting, S. Anders, E. E. Fullerton, M. R. Scheinfein
and H. A. Padmore. Science 287: 1014(2000)
Schulthess, T. C. and W. H. Butler. Phys. Rev. Lett. 81: 4516(1998)
Schulthess, T. C., W. H. Butler, G. M. Stochs, S. Maat and G. J. Mankey.
J. Appl. Phys. 15: 4842( 1999)
Schulz, H. J. Phys. Rev. Lett. 65: 2462(1990)
Skriver, H. L. The LMTO Method. Springer, Belrin (1984)
Smith, J. H. and E. R. Vance. J. Appl. Phys. 40: 4853(1969)
Soven, P. Phys. Rev. 156: 809 ( 1967)
Soven, P. Phys. Rev. 178: 1136 (1969)
Spisak, D. and J. Hafner. Phys. Rev. B 61: 11,569(2000)
Stohr, J., A. Scholl, T. J. Regan, S. Anders, J. Luning, M. R. Scheinfein,
H. A Padmore and R. L. White. Phys. Rev. Lett. 83: 1862 ( 1999)
Taylor, D. W. Phys. Rev. 156: 1017(1967)
Tomeno, I., H. N. Fuke, H. Iwasaki, M. Sahashi and Y. Tsunoda. J. Appl.
Phys. 86: 3853(1999)
Tsuchiya, Y., S. Li, M. Sano, T. Uesugi, S. Araki, H. Morita and M.
Matsuzaki. IEEE Trans. Magn. 36: 2557(2000)
Tsunoda, M., Y. Tsuchiya, M. Konoto and M. Takahashi. J. Magn. Magn.
Mater. 171: 29 (1997)
Umetsu, R., K. Fukam ichi and A. Sakuma. J. Magn. Magn. Mater. 239: 530
(2002a )
Umetsu, R. Y., K. Fukamichi and A. Sakuma. J. Appl. Phys. 91: 8873
(2002b)
Umetsu, R. Y., Y. Fujinaga and K. Fukamichi. J. Phys. Condens. Matter.
15: 4589 (2003)
Velicky, B., S. Kirkpatrick and H. Ehrenreich. Phys. Rev. 175: 747(1968)
von Barth, U. and H. Hedin. J. Phys. C 5: 1629(1972)
Wang, J. T., D. S. Wang and Y. Kawazoe. Appl. Phys. Lett. 79: 1507
(2001)
Wang, S. Q., W. E. Evenson and J. R. Schrieffer. Phys. Rev. Lett. 23: 92
( 1969)
Xi, H.and R.M. White. IEEE Trans. Magn. 36: 2635(2000)
Yamada, T., N. Kunitomi, Y. Nakai, D. E. Cox and G. Shirane. J. Phys.
Soc. Jpn. 28: 615(1970)
Yamaoka, T., M. Mekata and H. Takaki. J. Phys. Soc. Jpn. 31: 301(1971)
Yamaoka, T. J. Phys. Soc. Jpn. 36: 445(1974)
Yamaoka, T., M. Mekata and H. Takaki. J. Phys. Soc. Jpn. 36: 438(1974)
Antiferromagnetism of Mn Alloys 601

Yamauchi, R., K. Fukamichi, H. Yamauchi and A. Sakuma. J. Alloys


Compds. 279: 93( 1998)
Yamauchi, R., K. Fukamichi, H. Yamauchi, A. Sakuma.andJ. Echigoya. J.
Appl. Phys. 85: 4741(1999)
Yamauchi, R., T. Hori, M. Miyakawa and K. Fukamichi. J. Alloys Compds.
309: 16 (2000a)
Yamauchi, R., M. Miyakawa K. Sasao and K. Fukamichi. J. Alloys Compds.
311: 124(2000b)

The authors are grateful to Dr. C. Mitsumata of Hitachi Metals Ltd., Dr. R. Y. Umetsu and
Dr. K. Sasao of Tohoku University for many collaborations and discussions. The study of Mn
alloy systems was supported in part by a Grant-in-Aid for Scientific Research (8) (2), No.
13450255, from the Japan Society for the Promotion of Science.
Index

adiabatic temperature change 362,363 354,559,576,582,585


amorphous alloys 152 - 154, 167, 172, domain boundary 17,382,385,391
179,188,192,195,200,203,212,225, domain wall 10,12,16,17,21,24,36-
232,236,244,247 - 249,253,273,274, 42,44,51 - 54,57,59,62, 104,289,304,
279,280,283,302 382 - 385,390,393 - 395,418,434,436,
analytical electron microscopy 26,62 450,453, 455, 458, 460, 461 , 470, 488,
anhysteretic magnetization 392 - 394, 489,508, 509, 512, 519, 520, 522 - 524,
396,397 526 - 530,535,537,594,596
anisotropy energy 81, 393, 397, 399, EELS 29,30,32,33,144-146
421,422,427,475,499,527 elastic magnetic scattering 68
atom probe 267 - 270, 273, 274, 278 , electron holography 24,28,33,44,47,
280,282,284,285,287,290,300,304,306 52,56,57,63,65,130
atomistic scale 375,377 ,403 energy minimization 443,446,449,450,
CBED 32,121,122,124,145 453,477,495,496
CEMS 154,163,232 - 244 entropy change 358,359,361 - 364
crystallization 126, 151 - 154, 171 - exchange energy 422, 429, 430, 433,
175,177-183,185,188-193,195-197, 434,579,595
200 - 203,206 - 208,211 ,215,218,219, fast multipole method 407, 408, 410,
221 ,224 - 226,228,230,232 - 236 ,238- 413,417
244,250, 253, 257 - 259, 266, 271 , 272, FINEMET 172,178,180,182,185,186,
274 - 280 , 282,285 - 287 , 289,290,292, 188,189,230,232,270,272,275,277 -
294,296,300 - 303 280,282,284,287,301
Curie temperature 152, 179 - 181 , 191 , first-order phase transition 361
196,198,200,202,206,207,219,243, Fresnel method 24,36,41,42,51
248, 287, 290, 311, 313, 321, 323, 329, giant magnetocaloric effect 358
333, 335, 336, 340, 341 , 344, 345, 348, giant magnetoresistance 151 ,353 ,541
350,351 ,353 - 355,357 - 359,361 ,364, giant magnetostriction 353,356,357
366,367,400,430,550 HITPERM 172,211,230,243,287-289
density of state 315,316,344,476,549, hologram 25,28,47-49,51,53,59
561 HRTEM 113,118,126,128,130,145,
diffraction 31 , 32, 41 - 43, 52, 71 , 80, 146
84,87 - 89,91 ,92,98, 100,108,113, 115, hydrogen absorption 95, 350, 351 , 353 ,
118,121,122,124 - 128, 130, 131,135, 359,361
136,138,140,142 - 145,150,152,175, hyperfine interactions 154 - 156, 160,
176, 178, 182, 200, 244, 245, 249 - 251 , 162,163,168,169,208,244
259,273,279,280,283,345 - 347,351, hysteresis 13, 19, 81, 171, 317, 330,
Index 603

333,342,353,360,365,372 - 374,378, 161,163,166,168,178,180,181,189,


380,382,385 - 391 ,393,394,396 - 398, 191,192,200,201,207,211,232,244,259
402 - 405,446,458,460,463,465 - 469, nanocomposite 29, 30, 266, 267, 289,
484,485,508,510 - 515,517 - 525,532, 290,292,294 - 299 , 301 ,303 - 306,411 ,
537,539 424,437,452,463,464,484,485,499,502
inelastic neutron scattering 559 nanocrystall ine alloys 151 - 154, 163,
irreversibility 361 ,523 169-172,177,182,188,190,192,201-
isomer shift 155, 156, 164 - 167, 190, 203,206, 210 - 212, 229, 230, 245, 252,
192,195,202,203,215,216,245 253,285 - 287 ,298
itinerant-electron metamagnetism 310, nanoparticles 90, 140, 200, 219, 251,
311,367 256 - 259 ,267 ,282 ,283 ,463 ,499
La ( Fe,_ x Six) 13 344, 346 - 348, 350, NANOPERM 172,178,182,187 - 191,
355,358,359,362,366 202, 206, 207, 211, 230, 233, 242, 245,
Landau coefficient 311 , 313, 314, 323, 279,287,301
325 - 327 ,332,343 neutron 66 - 78, 80, 81 , 83 - 104, 107,
Landau expansion 312,313, 324, 330 - 108,156,158,180,181,244,249,252,
333,342,343,348,350 253,275,344,559,576,585,593
Laves-phase 310 - 312, 315, 318, 320, numerical simulation 464,472
340,344,354 phase reconstruction 25
Lorentz force 24,34,35,44 polarization analysis 68,75,78,83,85-
Lorentz microscopy 24, 26, 27, 36, 37, 87,91,100,101
42,44,62,65 potential 28,49, 60, 70, 72, 77 , 82, 92,
magnetic anisotropy 8, 81 , 115, 152 - 115,131,135,136,383,385,414,415,
154,169,170,182 - 186,211,212,215, 417,424 - 426, 438, 439, 463, 467, 492,
217 - 223, 227, 228, 230, 231 , 252, 259, 493,508,512,518 - 520,522,525, 527,
454 528,530 - 537, 539, 543, 545, 546, 548,
magnetic domain 2, 10, 12 - 14, 16, 17, 550 - 553 ,560,564,581
24 - 28 ,36,42,44,52,54,60 - 63 ,90,94, Preisach model 372, 374, 385 - 390,
108,148,382,385,391,508,593,595 398,508, 512 - 514, 520, 522, 523, 528,
magnetic excitations 485 530,539
magnetic phase diagram 311 ,313,324, pressure effect 324,332,350,590,591
335,344,559 quadrupole splitting 155, 164 - 167,
magnetic scattering 68,69,71 - 73,77 , 203,216,233,240,244,245,247
78,80,81,84,89,92,99,103,108 quasistatic and dynamic
magnetic spl itting 166, 167, 169 remagnetization 454
magnetism of nanocrystall ine alloys 154 rf-collapse effect 168, 169
microalloy 266,267,290,291 ,294,299, rf-Mossbauer 154, 168 - 171 , 182, 185,
304,306 186,190,192,206,210,212,218,219,
micromagnetic simulation 407,408,428, 222,223,225,227 - 232,244,245,247,
436,463,504,506 249,252,257,259
microstructure 24, 29, 33, 42, 81 , 182, rf-sidebands effect 169,170
185,186,192,193,208,232,266,267,271 selected reflection imaging 138, 140,
- 273, 275, 278 - 280, 282, 286 - 290, 143
292,294 - 299,301 ,303 - 306,378,390, short range order 89, 108, 154, 166,
505 180,194,203,216,221,244,247,249,273
Mossbauer spectroscopy 152 - 155, - 275
604 Index

specific heat coefficient 310,337-339, super-resolution TEM 130


588 thermal expansion 151 , 327, 329, 347,
spin fluctuation 104, 310 - 312, 318, 351,356,589,590,592,593
323, 324, 326, 327, 336, 338, 345, 353, thermal relaxation 508,509 ,536
365,367,593 thin films 9,26,48,57,66,72,90,95,
spontaneous volume 108,114,115,120,138,426,427,437,
magnetostriction 327, 345, 347, 351 , 450,460,528
590,593 transmission electron microscopy 24,
STEM 24,26,44,65,142 - 144 26,104,149,150,152,180,182,187,200
supermirrors 71 ,73,75 Triple-axis Spectroscopy 97
ERRATUM

Chapter 3, Characterization of Magnetic Materials by Means of


Polarized Neutron Scattering, G. Ehlers and F. Klose

Due to a mistake in the editing process, the revised version of Chapter


3 has not been included in this volume. The published chapter contains
typesetting errors and the list of references is incomplete.

The final version of this chapter can be obtained directly from the
authors or at the following uri:

http://w~vw.sns.gov/qQf!Hnentation/te5?b.p_lIbs.htm

Georg Ehlers (chlcrsg@ornl.gov)

Frank Klose (klosefr@ornl.gov)


Handbook of Advanced
Magnetic Materials
Volume III: Advanced Magnetic Materials:
Fabrication and Processing
Handbook of Advanced
Magnetic Materials
Volume III: Advanced Magnetic Materials:
Fabrication and Processing

Edited by:

Yi Liu
Center for Materials Research and Analysis
University of Nebraska
Lincoln, Nebraska

David J. Sellmyer
Center for Materials Research and Analysis
University of Nebraska
Lincoln, Nebraska

Daisuke Shindo
~1aterials
Institute of Multidisciplinary Research for Advanced Materials
Tohoku University
Sendai, Japan

(A)
(@) Tsinghua University Press ~ Springer
Library of Congress Cataloging-in-Publication Data

ISBN-IO: 1-4020-7983-4 e-ISBN-IO: 1-4020-7984-2


ISBN-13: 978-1402-07983-2 e-ISBN-13: 978-1402-07984-9

2006 Springer Science+Business Media, Inc.


All rights reserved. This work may not be translated or copied in whole or in part without the
written permission of the publisher (Springer Science+Business Media, Inc., 233 Spring Street, New
York, NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis.
Use in connection with any form of information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter developed is
forbidden.
The use in this publication of trade names, trademarks, service marks and similar terms, even if they
are not identified as such, is not to be taken as an expression of opinion as to whether or not they are
subject to proprietary rights.

Printed in the United States of America.

9 8 76 54 3 2 I SPIN 11097730

springeronline.com
Handbook of Advanced Magnetic Materials
Preface

In December 2002, the world's first commercial magnetic levitation supertrain


went into operation in Shanghai. The train is held just above the rails by
magnetic levitation (maglev) and can travel at a speed of 400 km/hr
completing the 30km journey from the city to the airport in minutes. Now
consumers are enjoying 50 GB hard drives compared to 0.5 GB hard drives ten
years ago. Achievements in magnetic materials research have made dreams of
a few decades ago reality. The objective of this book is to provide a
comprehensive review of recent progress in magnetic materials research. The
whole book consists of four volumes, each volume focusing on a specific field.
Graduate students and professional researchers are targeted as the readers.
Each chapter will have an introduction to give a clear definition of basic and
important concepts of the topic. The details of the topic are then elucidated
theoretically and experimentally. New ideas for further advancement are then
discussed. Sufficient references are also included for those who wish to read
the original work. Many of the authors are well known senior scientists. We
have also chosen some accomplished young scientists to provide reviews on
new and active topics.
In the last decade, one of the most significant thrust areas of materials
research has been nanostructured magnetic materials. There are several
critical sizes that control the behavior of a magnetic material. For example,
the coercivity of a magnetic material made of particles increases with
decreasing particle size, reaching a maximum where coherent rotation of a
single-domain particle is realized, and then decreases with further decrease of
the particle size. For a composite made of a magnetically hard phase and soft
phase, when the grain size of the soft phase is sufficiently large, the soft and
hard phases reverse independently. However, when the grain size of the soft
phase is reduced to a size of about twice the domain wall thickness of the hard
VI Preface

phase, the soft and hard phases will be exchange-coupled and behave as if a
single magnetic phase is present. Such behavior can be used to increase the
energy product of high-performance permanent magnets. Size effects become
critical when dimensions approach a few nanometers, where quantum
phenomena appear. The first volume of the book has therefore been devoted
to the recent development of nanostructured magnetic materials, emphasizing
size effects.
Our understanding of magnetism has advanced with the establishment of
the theory of atomic magnetic moments and itinerant magnetism. In general,
the magnetism of a bulk material can be considered as the superposition of
atomic magnetic moments plus itinerant magnetism due to conduction
electrons. In practical applications the situation becomes much more
complicated. The boundary conditions have to be taken into account. This
includes the size of the crystals, second-phase effects and intrinsic properties
of each phase. The effects of magnetic relaxation over long periods of time
can be critical to understanding. Simulation is a powerful tool for exploration
and explanation of properties of various magnetic materials. Simulation also
provides insight for further development of new materials. Naturally, before
any simulation can be started, a model must be constructed. This requires that
the material be well characterized. Therefore the second volume of the book
provides a comprehensive review of both experimental methods and simulation
techniques for the characterization of magnetic materials. After an introduction, each
section gives a detailed description of the method and the following sections
provide examples and results of the method. Finally further development of the
method will be discussed.
The success of each type of magnetic material depends on its properties
and cost which are directly related to its fabrication process. Processing of a
material can be critical for development of artificial materials such as
multilayer films, clusters, etc. Moreover, cost-effective processing usually
determines whether a material can be commercialized. In recent years
processing of materials has continuously evolved from improvement of
traditional methods to more sophisticated and novel methods. The objective of
the third volume of the book is to provide a comprehensive review of recent
developments in processing of advanced magnetic materials. Each chapter will
have an introduction and a section to provide a detailed description of the
processing method. The following sections give detailed descriptions of the
processing, properties and applications of the relevant materials. Finally the
potential and limitation of the processing method will be discussed.
The properties of a magnetic material can be characterized by intrinsic
Preface vn
properties such as anisotropy, saturation magnetization and extrinsic
properties such as coercivity. The properties of a magnetic material can be
affected by its chemical composition and processing route. With the continuous
search for new materials and invention of new processing routes, magnetic
properties of materials cover a wide spectrum of soft magnetic materials, hard
magnetic materials, recording materials, sensor materials and others. The
objective of the fourth volume of this book is to provide a comprehensive
review of recent development of various magnetic materials and their
applications. Each chapter will have an introduction of the materials and the
principals of their applications. The following sections give a detailed
description of the processing, properties and applications. Finally the
potential and limitation of the materials will be discussed.
NASA is considering the launching of spacecraft by maglev. The first
stage rocket, which accounts for two-thirds of the cost and is lost every
launch, would be replaced by a maglev track. Using a 50 ft track NASA
scientists have accelerated a model spacecraft to 96kph in less than half a
second. In the last few decades the knowledge of mankind has been expanding
rapidly into deep space measured by light years and the nano world where
building blocks of atoms are being engineered. Magnetism and magnetic
materials are among the most intriguing and fascinating science and
engineering fields. Undoubtedly advances in magnetic materials research will
continue to fuel our understanding of the universe in the new century. We hope
this book will provide a useful reference for researchers working at the frontier
of magnetic materials research.
We would like to express our sincere thanks to all our devoted authors,
technical editors, and publishers for making this book possible.

The editors
Contents

Preface V
List of Contributors -m
'i:V
1 HDDR Process for the Production of High Performance Rare-Earth
Magnets 1
11.. 1 Introduction '" 1
1.22 HDDR Phenomena in Nd-Fe-B
1. 3
1.2. 1 Hydrogen Absorption 4
1. 2.2 Disproportionation in Nd2Fel4
1.2.2 Fe14 B 4
1. 2.3 Recombination of Disproportionated Nd2Fe14 B
B 8
1. 3 Anisotropy in HDDR Treated Nd-Fe-B Magnets ... ... ...
1.3 11
1. 4 HDDR Phenomena in Other Rare-Earth Magnetic Materials
1.4 21
1. 4. 1 Sm2 Fel7
Fe17 Nx Compound 21
ThMnlrtype Compound
1. 4.2 ThMn12-type 23
1.4.3 Nd3 (Fe, Tj)29-Type
Ti)29-Type Compound 24
1. 5 Summary'" 25
References ... ... 25

2 Process and Magnetic Properties of Rare-Earth Bonded Magnets 32


2. 1 Introduction'"
Introduction ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 32
2. 2 Recent Progress on Rare-Earth Magnetic Powders for
Bonded Magnets 34
2. 2. 1 Isotropic NdFeB Powders 34
2.2.2 Lean Neo Isotropic Powders 38
2. 2. 3 Anisotropic NdFeB Powders for Bonded Magnets
2.2. 39
2. 2. 4 SmCo Powder for Bonded Magnet Applications
2.2.4 40
2.2.5 Other Magnetic Powders 43
2.
2.33 Process and Magnetic Properties of Polymer Bonded Rare
Earth Magnets 44
2. 3. 1 Bonded Magnets Produced by Compression Molding
Process 44
2. 3. 2 Hot Compression Molding Process
2.3.2 49
2. 3. 3 Bonded Magnets Produced by Injection Molding
2.3.3
Process 52
2. 3.4 Bonded Magnets Produced by Extrusion Process 53
X Contents

2.4 Polymers and Polymer Additives for Bonded Rare-Earth


Magnets 55
2. 5 Environmental Stability of Polymer Bonded Magnets
2.5 57
2 . 6 Summary 61
References 65

3 Laser Processing of Magnetic Materials 69


3. 1 Introduction'"
Introduction ... ... ... ... ... ... ... ... '" 69
3.2 Laser Characteristics and Laser Facilities for Materials
Processing ... ... ... ... ... ... ... ... 70
3. 2. 1 Laser Characteristics 70
3.2.2 Lasers Available for Materials Processing 71
3. 2. 3 Laser Materials Interaction
3.2. 74
3. 3 Laser Surface Treatment of Magnetic Thin Films 74
3.3. 1 Precise Grain Size Control of Magnetic Thin Films 75
3. 3.2 Alignment of Magnetic Grains by Laser Annealing
3.3.2 Annealing 78
3 . 4 Interferometric Laser Lithography
3.4 79
3. 4. 1 Multi-Step ILL '" ,. 80
3.4.2 Direct Patterning Using ILL , 82
3 . 5 Concluding Remarks 85
References 85

4 Processing and Properties of Nanocomposite Nd 2 Fe l4 B-Based


NdzFel4B-Based
Permanent Magnets '" '" , 88
Introduction
4. 1 Introduction'" '" '" '" ...
,..,. 88
4.2 Basic Considerations on Processing Techniques , 88
4 . 3 Preparation of Amorphous or Nanostructured Precursors
4. 91
4.4 Transformation of Amorphous or Nanostructured Precursors ... 97
4. 5 Direct Preparation of Nanocomposite Permanent Magnets'"
4.5 Magnets '" 99
4 . 6 Effects of Additive Elements on Crystallization
4.6 Crystal! ization Behavior 100
4.6.1 Fe 3B/NdzFe 14 B
Fe3B/Nd2Fe,4B 101
4.6.2 Advanced Nanocomposite Permanent Magnets
Based on Fe3B/NdzFe14B
Fe3B/Nd2Fe14B 103
4.6.3 cx-Fe/Nd
o:-Fe/Ndz2Fe14 B '" ,. , 104
4.7 Practical Properties of Nanocomposite Nd2
zFe14 B-Based
Magnets 106
4. 7. 1 Magnetic Properties 106
4.7.2 Corrosion Behaviors , 107
4.7.3 Magnetizability '" 108
4.. 8 Application
4 Appl ication Examples 109
4. 9 Future Subjects
4.9 110
References ... ... ... ... 111
Contents
COntents XI

5 Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring of


Magnetic Properties, Magnetoelastic and Transport Properties 115
5. 1 Introduction ..... ...... ...... ...... ...... ...... ...... ...... ...... ...... ...... ...... ...... ...... ...... ...... ......
'" ...... 115
5.2 Magnetic Properties: Magnetization Process 120
5.2. 1 Amorphous Ribbons: Effect of the Composition and
Applied Stress 121
5.2.2 Amorphous Wires: Large Barkhausen Effect 123
5.2. 3 Nanocrystall
5.2.3 Nanocrystalline ine Materials: Coercivity 128
5. 3 Processing: Induced Magnetic Anisotropy
5.3 Anisotropy .. 132
5. 3. 1 Amorphous Metall ic Alloys 135
5.3.2 Nanocrystalline
5. 3.2 Nanocrystall ine Alloys .. Alloys .. .. .. 139
5.3.3
5. 3. 3 Induced Magnetic Anisotropy in Glass-Coated
Amorphous Microwires 142
5.4 Magnetostriction ... ... ... ... ... 145
5. 4. 1 Amorphous Materials
5.4. 145
5.4.2 Nanocrystalline Materials 151
5.5
5. 5 Giant Magneto Impedance Effect 154
5.6 Sensor Applications 165
5. 6. 1 Magnetometers Using Amorphous and Nanocrystall ine
Materials 166
5. 6.2
5.6.2
Magnetoelastic Sensors: Sensors Based on Magnetic
Bistability 167
5.6. 3 Sensors Based on Giant Magneto-Impedance
Effect 172
References ... ... ... ... 174

6 Nanogranular Layered Magnetic Films 182


6. 1 Introduction ... ... ... , , 182
6.2 Ion Beam Sputtering 184
6. 3 Samples Preparation and Characterization
6.3 185
6. 4 Structural Properties 187
6. 5
6.5 Magnetic Properties 189
6.5.
6. 5. 1 Nuclear Magnetic Resonance 189
6.5.2 Room Temperature Static Magnetization 193
5. 3 dc and ac SQUID Measurements
6. 5.3 194
5. 4 Ferromagnetic Resonance
6. 5.4 200
6.6 Transport and Magnetotransport Properties 202
6.6. 1 Electric Conduction in Zero Magnetic Field 203
6.6.2 Room Temperature Magnetoresistance 205
6.6.3 Influence of the Number of Bilayers on Resistive
and Magnetoresistive Properties ... ... ... ... ... ... 207
6.6.4 Temperature Dependence of Magnetoresistance 209
6. 7 Conclusions 212
References ... ... ... 213
XII
}(j( Contents

7 Monodisperse Ferromagnetic Metal Particles: Synthesis by Chemical


Routes,
Routes. Size Control and Magnetic Characterizations 217
7. 1 Introduction 217
7.2 Preparation of Monodisperse Fine Ferromagnetic Metal
Particles by Chemical Routes 218
7.2. 1 Mechanism of Particle Formation 219
7.2.2 Control of the Morphological Characteristics of the
Particles ...... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 220
7.33 Characterization of Ferromagnetic Metal Particles Made
7.
by the Polyol Process ...... ... ... ......
... ... ... ......
... ... 226
7. 3. 1 Morphological Characteristics 226
7. 3.2 Crystal Structure 232
7.3.3 Chemical Composition 234
7.3. 4 Core-Shell Microstructure
7.3.4 237
7.3.5 Static Magnetic Properties 242
7.4 Magnetic Resonance in Fine Particles 248
7.4. 1 Theoretical Background 248
7. 4.2 Magnetic Resonance in Polyol Made Particles............ Particles.. 252
7. 5 Conclusions and Perspectives 261
References ... ... ... ... ... ... ... ... ... ... 262

8 Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation


and Characterization 267
8. 1 Introduction ... 267
8.2 The Half-Metallic Semi-Heusler NiMnSb 268
8.2.
8. 2. 1 Heusler Alloys 268
8.2.2 The Half-Metals 270
8.2.3 The Case of NiMnSb 272
8.3 Growth and Structure of Epitaxial NiMnSb Thin Films 275
8.3.1
8. 3. 1 PtMnSb
PtMnSb.. ... ... ... ... ... ... ... ... 275
8.3.2
8. 3.2 PtMnSb/NiMnSb Multilayers 275
8.3.3 Polycrystalline NiMnSb Thin Films 276
8.3.4 Epitaxial Growth of NiMnSb 279
8. 3.5 Surface Morphology of the Epitaxial Films
8.3.5 282
8.4 Magnetism and Transport Properties of NiMnSb Thin Films 286
8.4. 1 Hysteresis Cycles ...... ... ... 286
8. 4.2 Magnetocrystall
8.4.2 Magnetocrystalline ine Anisotropy 287
8. 4. 3 NiMnSb Thin Film Resistivity
8.4.3 .. 289
8.5 Evidences of the Half-Metallic Character 290
8.5. 1 Overview of Magnetic Excitations in Half-Metals........
Half-Metals 290
8.5.2 Electrical Resistivity 292
8.5.3
8. 5.3 Magnetoresistance 293
8. 5. 4 Hall Effect ... 294
8.5.5 Magnetization 297
Contents :xm
.xm

8.6
8. 6 Conclusions '" '" 298
References ... ... ... 298

9 Bulk Amorphous Magnetic Materials '" ,. '" 303


9. 1 Introduction '"
'" '" 303
9.2 Processing and Characterization of Bulk Amorphous Magnetic
Alloys 306
9.2. 1 Melt-Spinning Technique Used for the Preparation of
"Bulk" Amorphous Ribbons 306
9.2.2 Preparation of Bulk Amorphous Magnetic Rods and Rings
by Casting Techniques 308
9.2.3 Methods and Techniques Used for Structural and
Magnetic Characterization of Bulk Amorphous
Magnetic Alloys .. ..., ... ... ... ... , .. , 308
9.3 Soft Magnetic Bulk Amorphous Alloys 313
... ..., '"
9. 3. 1 Background'" ..... ... '"
... ... ... 313
9. 3.2
3.2 The Formation of the Amorphous Phase in
(Fe, Co, ND70(Zr,
Ni)70(Zr, Nb, M)lo M)lO Blo(M=Zr,
B20 (M=Zr, Ti, Ta, Mo)
Systems ... ... ... ... ..., ..
'" ........ ..., ... ... ... ... ... ... ... ... ... ... ... ... ... 314
9.3.3
9. 3. 3 Electric Resistivity of Fe-Based Bulk Amorphous
Alloys 317
9.3.4 Magnetic Properties of (Fe, Co, ND 70 (Zr, Nb, M) 10
Ni)70 lo Blo
20
Bulk Amorphous Alloys in the As-quenched State ... ... 319
9. 3.5 Magnetic Properties of Fe-Zr-B-Based Bulk
Amorphous Alloys Versus Annealing 321
9.4 High-Coercivity Bulk Amorphous Alloys 324
9.4. 1 Background'" ... ... ... ... ... ... ... , 324
9.4.2 Glass-Forming Ability in Nd-Fe-(AI, SDSystems
Nd-Fe-CAI, Si)Systems 325
9. 4. 3X-ray and Neutron Diffraction Experiments on NdgoNdgo - x
Aho-
Fe x A1 ySiyy Bulk Metallic Alloys
1o - ySi 327
9.4.4 Magnetic Properties of Ndgo-xFexAlto-ySiy
Ndgo-xFexAllO-ySiy Bulk
Amorphous Alloys '" '" '" 332
Mossbauer Spectra
9.4.5 M6ssbauer 342
9.5 Relationship between Microstructure and Magnetic Properties
in Bulk Amorphous Alloys 344
9.5.
9. 5. 1 Soft Magnetic Bulk Amorphous Alloys 344
9.5.2 High-Coercivity Bulk Amorphous Alloys 345
9.6 Future Perspectives 349
References 349

Index 353
List of Contributors

1. Satoshi Sugimoto Department of Materials Science, Graduate School


David Book of Engineering
Tohoku University
Aoba-yama 02, Sendai 980-8579, JAPAN
sugimots@material. tohoku. ac. jp
2. Jinfang Liu Electron Energy Corporation
Michael Walmer 924 Links Ave, Landisville, PA 17538, USA
jfl@electronenergy.. com
jfl@electronenergy
3. Yi Liu Center for Materials Research and Analysis
University of Nebraska, Lincoln, NE 68588-0113,
USA
Yliu@unlserve. unl.
un!. edu
Min Zheng MMC Technology
2001 Fortune Dr.
San Jose, CA 95131, USA
D. J. SeHmyer
Sellmyer Center for Materials Research and Analysis and
Department of Physics and Astronomy, University
of Nebraska at Lincoln, NE 68588-0113, USA
J. Mazumder 2141 GGB, 2350 Hayward Street
Department of Mechanical Engineering and Applied
Mechanics
University of Michigan
Ann Arbor, MI 48109-2125
4. Satoshi Hirosawa Research and Development Center, Sumitomo
Special Metals Co., Ltd. 2-15-17 Egawa, Shi-
mamoto-cho, Mishima-gun, Osaka-fu 618-0013,
Japan
HIROSAWA@ssmc. co. jp
5. Arcady Zhukov Instituto de Ciencia de Materiales, CSIC. 28049
Cantoblanco (Madrid). Spain
"TAMag Iberica S. L", Avda de los Remedios 41-
3A, Colmenar Viejo, 28770, Madrid, Spain
wupzhuka@scox01.sc.ehu.es
Julian Gonzalez Department of Materials Physics, Chemistry Facul-
ty, Basque Country University, 20018, San
Sebastian, Spain
XVI List of Contributors

6. G. N. KaKazei Department of Physics, Ohio State University, 174


West 18th Ave. , Columbus, OH 43210, USA
Yu. G. Pogoreloy CFP, Universidade do Porto, Rua do Campo Ale-
gre 687, 4169-007 Porto, Portugal
J. B. Sousa IFIMUP, Faculdade de Ciencias,
Cielncias, Unversidade do
J. M. Santos Porto, Rua do Campo Alegre 687,
4169-007 Porto, Portugal
S. Cardoso INESC, Rua Alves Redol 9-1, 1000-029 Lisbon,
P. P. Freitas Portugal
7. Philippe Toneguzzo CEA Le Ripault
BP16
F-37260 Monts, France
Tel: + 33247344897;
33247344897 ;
Fax: + 33247345179;
+33247345179;
toneguzzop@ripault. cea. fr
Guillaume Viau Universite Paris 7 Denis Diderot
Fernand
Femand Fievet Laboratoire de Chimie des Materiaux Divises et
Catalyse
2 Place Jussieu
F-75251 Paris Cedex 05, France
Tel: 33144277698;
Fax: 33144276137
viau@ccr.jussieu.
viau@ccr. jussieu. fr
Tel: 33144272805;
Fax: 33144276137
fievet@ccr. jussieu. fr
fievet@ccr.jussieu.
8. Delia Ristoiu CNRS, Laboratoire Louis Neel
J. -Po Nozieres 25 Avenue des Martyrs, B. P. 166, 38042 Greno-
ble, France
ristoiu@labs. polycnrs-gre. fr,
nozieres@phsmems.com
9. H. Chiriac National Institute of Research and Development for
N.Lupu Technical Physics
47 Mangeron Boulevard
6600 lasi, Romania
1 HDDR Process for the Production of High
Performance Rare-Earth Magnets

Satoshi Sugimoto and David Book

In this chapter the term "magnetic polarization" J is used, which is related to


"magnetization" M and "magnetic flux density" B, by J = 1J0 1-10 M = B - 1J0
1-10 H,
where magnetic constant (permeability of a vacuum) 1J0 = 4n
1-10 = 4lT x 7
X 10- (H/m). In
general, SI units were used instead of cgs units. However, the symbols and
units in the figures were not changed from those used in the original papers,
and so cgs units sometimes remain. In addition, for the units of magnetic field,
both amperes/metre [Aim]
[AlmJ (related to H) and Tesla [T] [TJ (related to 1J0
1-10 H)
were used.

1. 1 Introduction

Permanent magnets are now essential components in many fields of


technology, because of their abi ability
Iity to provide a magnetic flux, and they have
applications in a wide range of devices (Fastenau and van Loenen, 1996). In
the last few decades, the development of hard magnetic materials has been
very rapid, with the advent of rare-earth (R) permanent magnets.
In the 1960s, a hexagonal SmCos compound with the CaCus-type CaCus -type structure
appeared as the first rare-earth high performance magnet (Strnat et al. ,
1967; Strnat, 1988). The compound has magnetic properties that are suitable
for a permanent magnet, such as a large uniaxial magnetocrystalline
anisotropy 1J0 ..... 28 T (H A
1-10 H AA ....., A is anisotropy field [Aim]),[A/mJ), a relatively high
saturation magnetization (J ss"'" ....., 1. 14 n,
T), and a high Curie temperature (T c = =
720 C). The energy product, (BH) (BH ) max'
max , of this type of magnet reached

160 kJ/m 3 (20 MGOe). In an attempt to produce compounds with even higher
magnetizations, the Co content was increased to give R2Co CO 17
I7 .. The Sm2 C0
CO 17
l7
compound has a higher saturation magnetization (J s "'" 1. 25 n
Us""" T) and a higher
Curie temperature than the SmCos compound, however the anisotropy field is
smaller. In 1968, Nesbit et al. (1968) reported that Cu-added SmCos ingots
(i. e. , not processed by powder metallurgy) exhibited high coercivities (over
320 kA/m) after heat treatment, and Tawara and Senno (1968) reported a
similar effect for Cu in CeCos ingot alloys. These are the first reports of the
development of a two-phase decomposed R-Co magnets. A compositional
compromise was then developed, between the high magnetization of Sm2Col7 Sm2 CO l7
2 Satoshi Sugimoto and David Book

and the high magnetic hardness of SmC0 SmCo5s ', by a precipitation hardening
process, which forms Sm2 CO l7 phase surrounded by a SmC0 SmCo5s -type boundary
phase. This Sm2 C0 CO 17
l7 -type alloy, with additions of Fe, Cu, and Zr, increased
the maximum energy product to 264 kJ/m 03 3 (33 MGOe) (Mishra et al. , 1981).
The development of RCo-based permanent magnets has been reviewed in
detail by Strnat and Strnat (1991).
The main disadvantage with these materials was the comparatively high
cost of the Sm and Co, and so research began to focus on trying to find an Fe-
based magnetic material, with similar properties. This eventually led to the
joint announcement of the development of magnets based on the body centred
tetragonal Nd2Fe14 B phase, by Croat et al. (1984a, b) using melt spinning,
and Sagawa et al. (1984) using powder-metallurgy techniques which resulted
in energy products greater than 288kJ/m 3 (36 MGOe). The magnetic
characteristics of the new Nd2Fel4 Fe14 B phase are J s5 = 1.61 T, J.1o
Po H A = 7 .2 T and
T c = 312 'C. Due to their superior magnetic properties and low cost, the Nd-
Fe-B magnets have rapidly replaced SmCo type magnets (with the exception
appl ications, due to the relatively low T cc of the Nd2Fel4B
of high temperature applications, Nd2Fe14 B
phase). Since then, Nd-Fe-B magnets with a (BH)max over 400 kJ/m 3 have
been produced commercially, by improving alloy and powder preparation,
magnetic pressing, and surface coating (Herbst, 1991). Very recently, an energy
product of 460 kJ/m3(57.6 MGOe) was reported by Kaneko (2000,2004).
The search for novel and improved hard magnets has continued, with the
discovery of a number of promising magnetic materials such as the ThMnlr ThMn12-
type compounds (Ohashi et al., al. , 1987; Mooij and Buschow, 1987), Sm2Fe17Sm2 Fel7
interstitially modified with nitrogen (Coey and Sun, 1990), ThMnlrtype
(Yang, 1991) and Nd3(Fe,Ti)29-type (Collocott et aI., 1992, Cadogan et
al. , 1994) compounds. The most notable of these compounds is Sm2 Fel7 Fe17 Nx ,
which offers the prospect of magnets with even better magnetic properties and
a high Curie temperature (476 'C), compared to that of Nd2Fel4 Fe14 B
(Kobayashi, 1994; FujiiFuj ii and Sun, 1995; Skomski,
Skomsk i, 1996).
There are two well-established techniques for the manufacture of rare
earth permanent magnets: powder metallurgy is used to obtain high
performance, anisotropic, fully dense magnet bodies; and melt-spinning is
widely used to produce magnet powders for isotropic bonded magnets.
Although the magnetic properties (i. e., the energy product) of near fully
dense sintered magnets are superior, the ability to fabricate components to
near net shape using coercive powder is often a more important factor (and
which according to Fastenau (1996) and Luo (1999), has lead to bonded Nd-
Fe-B magnets becoming the most rapidly growing sector of the permanent
magnet market)
market> .
A more recent technique is the Hydrogenation, Disproportionation,
Desorption, and Recombination (HDDR) process, which consists of a series of
heat treatments in hydrogen and under vacuum. HDDR has proved to be an
effective and economic way of obtaining powders for use in the production of
HDDR Process for the Production of High Performance. . . 3

high performance, anisotropic bonded magnets.


There have been a number of articles that have reviewed aspects of the
HDDR process (Harris, 1996; Gutfleisch and Harris, 1996; Buschow, 1997;
Gutfleisch and Harris, 1998).
1998), In this chapter, we review the history of the
HDDR process, including recent improvements in the processing conditions
used for the production of anisotropic Nd-Fe-B HDDR powders, and HDDR
phenomena in other rare-earth iron (R-Fe) based compounds.

1. 2 HDDR Phenomena III Nd-Fe-B

The HDDR process was first reported in Nd-Fe-B magnets by Takeshita and
Nakayama (1989, 1990), and the reaction mechanisms involved were clarified
by McGuiness et al. (1990a, b) and Harris and McGuiness (1990). A
schematic illustrationof the conventional HDDR process is shown in Fig. 1. 1.
In this process, the Nd-Fe-B ingot is heated to '" 700 - 900 C (which is often
accompanied by the decrepitation of the ingot into powder) and kept at this
temperature under hydrogen, and then heat-treated under vacuum. After
cool ing under vacuum, a coercive powder is obtained which can be used to
form bonded magnets. This process consists of four steps: hydrogenation of
Nd2z Fe14 B; disproportionation of Nd2z Fe14 B into NdH
NdH2+5' Fe2 B;
z+8' ex-Fe, and Fez
desorption of hydrogen gas from NdH2+5;
NdH z+8; and finally, recombination to the
Nd2zFe14 B phase. During this process grain refinement occurs, resulting in a
,-----------{
, - - - - - - - - { Treatment temp.:700-950 '('
'C ))---------,
------,

Recombination stage

Conventional
HDDR treatment

H
He2 Evac.
111,=0.1
PH,=O.I MPa

Figure 1. 1 Schematic illustration of the HDDR process in Nd-Fe-B.


Nd-Fe-B,
4 Satoshi Sugimoto and David Book

coercive powder that contains Nd2Fe14 B grains similar in size ((,,-, O. 3 IJm) to
..... 0.3
the critical single-domain size for the Nd2Fe14 B phase (0.24 IJm) .

1. 2. 1 Hydrogen Absorption

Shortly after the discovery of Nd2Fel4Fe14 B, the properties of the hydride began to
be widely researched. I' Hertier et al. (1984) showed that hydrogen
absorption into the Nd2Fe14 B phase leads to increases in the unit cell volume
( .....
"-' 4 % ), saturation magnetization (,,-,( ..... 6 %) and Curie temperature (about
80 'c higher). However, there is also a remarkable decrease in the
80C
anisotropy field, resulting in a very low intrinsic coercivity (Wiesinger et a!.al. ,
1987). These changes in magnetic properties can be correlated to an increase
in the Fe-Fe distances within the hydride crystal structure (Isnard et al. ,
1995)
1995)..
Following on from earlier work (reviewed in Harris, 1987) on
SmCos and Sm2(Co,Fe,Cu,Zr)wtype
Sm2 (Co, Fe, Cu, Zr) 17 -type magnet alloys, Harris et al. (1985)
demonstrated that the hydrogen decrepitation (HD) process can also be
applied to Nd-Fe-B alloys. Exposing the Nd-Fe-B cast alloy to O. 1 MPa of
hydrogen reduces the ingot into a friable powder that is much easier to mill
than mechanically crushed ingot. Decrepitation occurs due to large volume
expansions during hydrogen absorption, in certain phases: 4.8 % for
Nd2Fe14 BH2.7' and 16.4 % for NdH 3 (Oesterreicher and Oesterreicher, 1984).
Studies have shown HD to be an effective, and economical, extra step in the
powder metallurgy route for producing sintered Nd-Fe-B permanent magnets
(Harris et al.a!. , 1995).
The temperature at which hydrogen is absorbed in Nd-Fe-B alloys is
strongly dependent upon the presence of the Nd-rich phase: in a near
stoichiometric Nd2 Fe14 B alloy, hydrogen absorption will only take place if
heated to at least 150 C 'c ,, compared to Nd-rich alloy, which can be hydrided
at room temperature after a few minutes (Harris et al., 1987). This is
because the Nd-rich phase has much lower activation energy with respect to
hydrogen than Nd2Fe14 B, and so it will absorb hydrogen at a lower
temperature. The heat generated by the exothermic hydrogenation of the Nd-
rich phase then activates hydrogen absorption into the Nd2Fe14 B phase (Harris
a!., 1987, Book and Harris, 1992, 1995).
et ai.,

1. 2. 2 Disproportionation in Nd z2 Fel4
Fe14 B

When Nd2Fe14 B is heated to temperatures 'c, X-ray diffraction


over 600 C,
(XRD) and thermomagnetic analysis (TMA) show that the NdNd2Fe14B
2Fe14 B phase
disproportionates into ex-Fe, Nd hydride and Fe2 B (I' Heritier et al., 1984;
a!. , 1985).
Harris et al.
HDDR Process for the Production of High Performance. ..
.. 5

Nd 2Fe 14 B
Nd2Fe14B + (2 x)H 22 ~ 12ex-Fe + 2NdH2x + Fe2B t:.H
llH (1.1)
where t:.H
llH is enthalpy change (J/mol). The range of the reaction was
investigated by Cadogan and Coey (1986) using a thermopiezic analyzer
(TPA) , which demonstrated that, for Nd 15 Fen Be, B8 , hydrogenation at ....., 'c
....... 200 C
was followed by a gradual desorption of hydrogen up to ....., 'c,, at which
....... 650 C
point a second rapid absorption occurred, finishing at ....., ....... 800 C. 'C. Mossbauer
spectroscopy (Cadogan and Coey, 1986; Yu et al., 1991) showed that the
second absorption corresponded to the disproportionation reaction, with ex-Fe,
Fe2 Band NdFe4 B4 being d~tected, while XRD confirmed the presence of Nd
hydride. Hydrogen absorption/desorption studies showed that the
disproportionation and recombination reaction temperatures depend on the
initial grain size of the alloy, with significantly lower temperatures for
materials with smaller grain-size, such as melt-spun ribbons (Book and Harris,
1992; book et al. 1995; Meisner and Panchanathan, 1993).
As discussed in Section 1. 2. 1, heating Nd-Fe-B alloys under hydrogen
will result in ingots decrepitating into a fine powder; the temperature at which
this occurs will depend on the amount of Nd-rich phase present in the alloy,
hydrogen pressure and the condition of the surface of the ingot. However,
Zhang et al.a!. (1991) showed that it is possible to avoid hydrogen decrepitation
by heating near-stoichiometric alloy very rapidly in hydrogen (>40 C 'c /min) ,
to obtain a sol id block of disproportionated (or, if subsequently exposed to a
vacuum, recombined) material. Alternatively, hydrogen can be introduced at
an elevated temperature, which has been termed solid-HDDR (Gutfleisch
a!., 1994).
et ai.,
Nakamura et al. (1995) also investigated the hydrogen absorption and
desorption characteristics of the Nd22Fe14 B compound by monitoring
quantitative changes in the flow of hydrogen entering (Oin) (Qin) and exiting (Oout) (Qout)
the heat-treatment furnace, as shown in Fig. 1. 1.2.
2. When the gas flow rate is
constant, the difference between Qauto out and OinQ in is equivalent to the quantity of
(llO = Q
gas absorbed into the sample (t:.Q Oout-
out - Qjn)' i. e. desorption occurs when
Oin),
t:. Q
II 0 is positive and absorption when it is negative (Nakamura et al., ai., 1992).
Furnace

Figure 1. 2 A schematic illustration of the gas-flow equipment used to measure the


hydrogen absorption and desorption characteristics (Nakamura et al. , 1992; Sugimoto et
al. , 1992).
6 Satoshi Sugimoto and David Book

Figure 1.3 shows the hydrogen absorption and desorption characteristics of a


Ndlz.6Fe81
Nd 60 alloy: upon heating in hydrogen, absorption occurs at 250 'c
4B60
12 . 6 Fe81 4B

which corresponds to the formation of Ndz 2 Fe14 BH x ' followed by a steady


desorption between 250 - 650 'c, and then an absorption peak at around
700 'c corresponding to the disproportionation reaction. Nakamura et al.
(1995) also studied the effect of additional elements on the disproportionation
reaction and found that the Zr and Nb additions made the disproportionation
reaction sluggish and shifted disproportionation to higher temperatures, as
shown in Fig. 1.4. 1. 4.

Nd12.6FeSI.4B6.0
Nd12.6FCSI.4B6.0 in H 2
Heating rate: 400C /h
Particle size: <63 ~m
20 Sample weight: 5 g
Flow rate: 200 ml/min

-~
- ........_ - - - - -
~ _~-----

-20

300 600 900


eCl
Temperature CC)
Figure 1. 3 Hydrogen absorption and desorption characteristics of Nd 12
l26 FeS1.8 B60
. 6 Fe8lS

on heating in a flow of hydrogen at a rate of 400 C


'O/h
/h (Nakamura et ai.,
al. , 1995).

Other in situ techniques that


that have been used to characterize the
disproportionation reaction, while heating Nd-Fe-B under hydrogen, include:
OTA (Scholz et al. , 1987; Book and Harris, 1995), thermomagnetic analysis
DTA
(Book and Harris, 1995; Burkhardt et al., 1998), electrical resistivity
(Gutfleisch and Harris, 1995a), and powder neutron diffraction analysis
(Lisert et a!.
al. , 1997).
After the discovery of the HDDR
HOOR process, further work was carried out to
characterize the nature of the disproportionated Ndz2 Fe14 B materials. Xiao et
al. (1992) showed by transmission electron microscopy (TEM) that the
a!.
disproportionated grains were about O. 1 IJm in diameter. Gutfleisch et al.
(1994, 1995b) have shown by TEM analysis that the disproportionated matrix
ex-Fe and that within
exhibits a colony-type structure consisting of NdH_ z2 and a-Fe
each colony rods of NdH_ z2 are embedded in an a-Fe
ex-Fe matrix. Itakura et al.
a!.
( 1998) and Nakamura et al. (1996) observed similar microstructures, and
showed that the morphology of the phases are affected by the
disproportionation temperature. At relatively low temperatures, rod-shaped
HDDR Process for the Production of High Performance. . . 7

Nd 12.6 FeS1.4-,Zr,B 6.0


NdI2.6FeSI.4-xZrxB6.0 in H 2
Heating rate: 400'C Ih/h
Pal1icle
Particle size: <63
<63/Jm
/lm --x=0
--x=O
Sample weight: 5 g _ . - x=O.1
20
Flow rate: 200 ml/min _.- - x=1.0
------

'2
] ~ ~ ~
I()) 0 ....\
<l
'V
-20 \1
~
I

600 300 900


Temperature CC)
('C)
Figure 1. 4 Hydrogen absorption and desorption characteristics of Nd 126
12 6 FeS14- x Zr x 8 60 on

heating in a flow of hydrogen at a rate of 400 'c /h, (Nakamura et ai,


"C /h. al. , 1995),
1995).

NdH_ 2 was observed, however, at higher temperatures, spheres of the NdH_ 2


phase were found: this is thought to occur in order to decrease the interfacial
energy.
energy, These microstructural changes are summarized schematically in Fig,
Fig. 1,
1.5.
5,

~Nd2FeI4B
~Nd2Fe14B
0(O-100/Jm)
W(D-l00/lm)

n
B Heat in H
He2

-600 C
0

::;;,
""'.:'
'iii'.
"ii'
,,;'!!':,
.,,'''''.
Oisproportionation
Disproportionation
reaction starts at the
grain boundary

n
Dispropol1ionation
Disproportionation
is complete

"
Fe
'-G"

B
~.


Spherical NdH2
2
embedded in Fe
Heat inH 2

High
High temperature
temperature
(-900 'C)
"C)

NdH 2
Figure 1. 5 Schematic illustration of microstructural changes during the disproportionation
reaction in a Nd-Fe-8 alloy, for the conventional HDDR treatment(
treatment(Adapted
Adapted from a figure in
Sugimoto et al
al. , 1999),
1999).
8 Satoshi Sugimoto and David Book

Nakamura et al. (1996) observed the following relationship in


crystallographic orientation between the decomposed mixture of NdH2
z and ex-
Fe:
<111
111>a_Fe
) a-Fe II <111>NdH
< 111 ) NdH2
2
'' and (110)a-Fe
(110) a-Fe II (110)NdH
(110) NdH2
2
..

Although higher temperatures cause the grain growth of these phases, the
crystallographic relationship is maintained. In addition, a relationship was
found between the disproportionated mixture and the undecomposed Nd2 zFe!4
Fe14 B
phase:
<111>
111 >a-Fe
a-Fe and <111 >NdH2
111>NdH 2
II <00 11>>Nd22 Fe"
Fe!4 B
((llO)a-Fe
110) a-Fe and (110)
(110)Nci-l
NdH
22
II (100) Nd Fe" B with a canting angle of several degrees.
(100)Nd2Fe!4B
2

However the crystallographic relationship between the undecomposed Nd2 z Fe'4


Fe14 B
phase and the disproportionated mixture only exists on a local scale.
Gutfleisch et al. (1994) reported the same relationship between the
decomposed NdHz2 and Fe.
In contrast to these reports, Tomida et al. (1996, 1997)1997> reported that no
apparent orientational coherency exists between the disproportionated
products of ex-Fe, FezFe2 Band NdHz2 and the parent Nd2 z Fe'4
Fe14 B by transmission
electron microscopy (TEM) and high resolution transmission electron
microscopy (HRTEM). However, they observed nanoscale Nd2 zFe'4
Fe14 B particles
of 10 to 100 nm in diameter, whose crystallographic axes were nearly parallel
to those of the original Ndz2 Fe'4
Fe14 B, present throughout the disproportionated
mixture. They suggested that these nano-scale Nd2 z Fe,4
Fe14 B particles act as
recombination centres during the desorption stage, leading to the formation of
anisotropic grains (described in the next section).
Recently Tomida et al. (1999) detected an intermediate hydrogenation
phase-tetragonal Fe3B Ct-Fe3
(t-Fe3 B) -in the early stage of disproportionation. The
t-Fe3 B phase possesses lattice coherency and associated "one-to-one"
orientation relationship with the parent Nd2 zFe'4
Fe14 B phase:
(001)1
(001), II (001)Z-14-1
(001)2-14-1
[100J1
[100J, II [100Jz-14-1
[100J2-14-1'.
Within the t-Fe3 B grain, nano-scale Ndz2Fe14 B particles exist along with NdH2
z
particles. Tomida et al. (1999) suggested that t-Fe3
t-Fe3BB may play an important
role in the anisotropy-mediating mechanism.

1. 2. 3 Recombination of Disproportionated Nd z2 Fe14 B

Oesterriecher (1985) postulated that it should be possible to recombine the


disproportionated mixture of ex-Fe, Nd hydride and Fez
Fe2 B, by heating under
vacuum at elevated temperatures to develop" microcrystals of Nd2
develop "microcrystals z Fe!4
Fe14 B. "
Subsequent attempts to recombine the disproportionated mixture back into the
HDDR Process for the Production of High Performance. . . 9

Nd2zFe 14B
Fe14 B phase (Harris et a!.
ai.,, 1985; Pollard and Oesterriecher, 1986) were
not successful. However, when Takeshita and Nakayama (1989) found that a
magnetically coercive powder could be produced by exposing Nd-Fe-B ingot to
a combination of hydrogen and vacuum heat treatments at 700 - 900 C 'c ,,
McGuiness et a!. al. (1990a, 1990b, and Harris and McGuiness, 1990)
interpreted this behavior in terms of a Hydrogenation, Disproportionation,
Desorption, Recombination (HDDR) reaction. The HDDR process can convert
coarse grained ingot materials into very fine grained powder capable of
exhibiting large intrinsic coercivities.
The standard HDDR treatment is to heat the alloy under hydrogen (at
which stage, ingots of Nd-Fe-B will decrepitate into powder, as discussed in
Section 1.2. 1), to a reaction temperature between 750 and 850 C 'c , where it
is kept for 2 h under hydrogen and 1 h under vacuum, and then cooled to room
temperature under vacuum. The reaction temperature is critical with respect to
coercivity. The presence of a maximum has been attributed, (Takeshita and
Nakayama, 1990; McGuiness et al., ai., 1990a, 1990b; Zhang et al., 1991)
1991> to
the establishment of an optimum microstructure around 0.3 IJm, such that in the
lower temperature regime, the coercivity is degraded by the presence of soft
ferromagnetic free iron, whereas at higher temperatures it is limited by the
presence of easily demagnetized large grains. Nakayama et a!. al. (1991)
observed the microstructure after the recombination reaction using TEM and
reported that the grain size of recombined grains was around 0.3 IJm and that
no grain boundary phase was found, as shown in Fig. 1.6.

Method Melt-spinning HDDR

Nd2Fe l4 B
10-2 10- 1 10 1
size(llm)
Boundary
Amorphous X
X Nd-rich
phase

Figure 1. 6 Structural comparisons of Nd-Fe-B magnets produced by melt-spinning,


metallurgy/ sintering (Nakayama et al. , 1991).
HDDR and powder metallurgy/sintering

Hydrogen absorption/desorption studies by Nakamura et al. (1995)


'c is due to recombination to
showed that a desorption peak at around 1000 C
the Nd2Fe14B
NdzFe14 B phase in hydrogen. An endothermic peak (and mass decrease)
can be seen at around the same temperature in the DTA (and TG) profile of
10 Satoshi Sugimoto and David Book

Nd 16 Fe76 B8a heated under hydrogen, in the paper by Scholz et al. (1987).
Nakamura et al. (1995) investigated the effects of additional elements
such as Co, Ga on the recombination temperature by using 6.0 flO
measurements, as shown in Fig. 1. 7. They reported that under O. 1 MPa of
hydrogen, the recombination reaction occurs at a temperature that is about
100 'c'C lower than the dissociation of NdH z2 ,, and Co or Ga added Ndz2 Fe14 B
alloys exhibit much lower recombination temperatures than the ternary alloy.
Therefore, they concluded that the dissociation of NdHz2 is not the rate-
determining step in the recombination reaction of Ndz2 Fe14 B, which instead
must be related to the free energy changes of formation of both the NdH z2 and
Nd2z Fe14 B phases. In addition the effect of the additional elements such as Co
and Ga is to change the recombination condition in the temperature and
hydrogen pressure. Sugimoto et al. (1997) measured the resistivity change
during the recombination reaction, and then produced TTT diagrams of the
disproportionation and recombination reactions, as shown in Fig. 1. 8. This
showed that the temperature and completion time of the recombination reaction
is decreased by Co or Ga addition. Sugimoto et al. (1997) also suggested
that the hydrogen pressure and temperature decrease by the additions, and
concluded that the main effect of additional elements is to change the HDDR
conditions.

:'::':'::':'~:::h;H~I7~1(6~O
Heating rate: 400C Ih
Particle ~\ :
20 Particle size:
size: <6311m
<63 1lm ."f 1'\
Sample weight: 5 g f.
Flow
Flow rate: 200 ml/min
rate: 200 mllmin ir\
~

---
c: Ii \
] ..'-'-- _ _ _ _ _
t . .
_....,.:::."..,_~~'b::......_
II 0 , : -=-~~--

~01
<] Ii --x=o
--x=O
f ___ x=11.6

-20
I _. x=17.4
------ x=17.4

300 600 900


TemperatureCC)
Temperature ("C)

Figure 1. 7 Hydrogen absorption and desorption characteristics of


Nd126Fes1.4-xCoxB60
Nd126FeS14-xCoxB6,O on heating in a flow of hydrogen at a rate of
400 C
'C /h (Nakamura et al.
aI.,, 1995).
HDDR Process for the Production of High Performance. . . 11

880
Disp.

~
860

840
f
? . . . .,P
.P

?? rr
~
h 820
i
::l 800
,3
9;0
i
E
780
\

~ 760 \j
~~
::

740
720 L..-
'---- ----!-
---,L --'-,-
-,L-- ~
~

1 10 100 1000
Time (disp), td(min)
Id(min)
(a)
880
......O .. .. Co(50%)
0Co(50%)
Rec.
860 --6--CoGa(compl.)
--.Ao--CoGa(compl.)
C(\<
C(\. .t ......t;.... CoGa(50%)
{;CoGa(50%)
~ 840 ..~? .....
~ 4:00 qq
h 820
: ..........
i
'"
::l

~0.
E780
800
l"',\9
i\\\\
~
~ 760 -ter.(compl.j'-..\.
f-" -ter.(compl.)"-.. \ \.\
740 .... 0.... ter.
.. 0 (50%) :h
ter.(50%) .... \ 1\
- C o (compl.) 0 0 n
720'------:-'::-------:-:'-::----..,..-:-'
720'----------,-1:,-------,--:':-----:-:-'.
1 10 1000
Time (rec), tr(min)
Time(rec),lr(min)
(b)

Figure 1. 8 TTT diagrams of (a) disproportionation and (b) recombination


reactions for Nd,26
Nd 12 .6Febal Nd,2.6
Feba, 8 60 , Nd CO I 1.0 8 60 and Nd
l26 Feba, 0011.0 Nd,2.6
12 .6 Febal
Feba, CO I 1.0 GalO
0011.0 Ga1.0 8 60
60
alloys, constructed using data from in situ resistivity measurements (Sugimoto
et al. , 1997).

1. 3 Anisotropy in HDDR Treated Nd-Fe-B Magnets

Takeshita and Nakayama (1992) showed that anisotropic powder could be


produced by the addition of very small quantities of certain elements.
Additions of Co were also found to introduce anisotropy, as shown by the basic
alloy in Fig. 1.9 denoted by an Fe "addition". Their studies indicated that the
elements Zr, Hf and Ga were the most effective for the production of
anisotropic magnet powders.
12 Satoshi Sugimoto and David Book

In magnetic field
8 No magnetic field

Al Si Ti V Cr Fe Cu
Cn Ga Ge Zr Nb Mo In Sn Sb Hf Ta W Pb Bi
l~
15






o
AI
n ~ r.l
Ie1 ra1
r.1

Al Si Ti V Cr Fe Cu Ga Ge Zr Nb Mo In Sn Sb Hf Ta W Pb Bi
n
Figure 1. 9 Magnetic properlies
properties of HDDR treated Nd-Fe-Co-B-M (where M is an
== Fe corresponds to Nd-Fe-Co-B) magnet powders. (Takeshita and
additive; M =
Nakayama, 1992).

Takeshita and Nakayama (1993) outl ined three possible explanations for
the "memory effect" responsible for anisotropy in their Nd-Fe-Co-B alloys:
(1) "A crystalline
crystal Iine direction, size, distribution and/or fluctuation of
composition within the transformed NdH2 , cx-(Fe,Co)
o:-(Fe,Co) or (Fe, CO)2B."
(2) "The transformation does not take place completely or proceeds
slowly, leaving a small amount of Nd2 (Fe, Co) 14 B phase that exists together
with NdH 2 , cx-(Fe,
o:-(Fe, Co), or (Fe, CO)2 CO)2B,B, and fine grains on the basis of this
Nd2(Fe,Co),4B
Nd2(Fe, Co) 14 B phase".
(3) "New phases made by additive elements".
However, TEM studies by the same authors, found no evidence for (3). (3) .
Harris (1992) suggested that additions such as Zr might stabilize parts of
Fel4 B phase with respect to disproportionation, and thus maintain an
the Nd2Fe'4
orientation relative to the original Nd2Fe'4 Fel4 B grains. On subsequent
recombination of the surrounding disproportionated material, these stabilized
regions would act as nucleation points for new grains, which would grow with
the same orientation as the original grains. This model gained support when
Buschow (1994) explained how such stabilized regions might be explained in
thermodynamic terms: in the case of Zr addition differences in the hydride
formation enthalpy for Zr2 Fe'4 Fel4 B (Capehart et al. 1993) and Nd 2Fel4 Fe14 B
suggested that Zr-stabilized regions would be less likely to disproportionate.
Uehara et al. (1993) reported that Nd2Fel4 B could be detected in
disproportionated NdNd,2
125.5 Fe70-
Fe70-xx CO'I 5 Ga x B6 alloys by XRD, and that as the
COil 5
amount of Ga was increased from x = 0 to 5, the proportion of Nd2Fel4 B phase
increased. They suggested that, for an Nd-Fe-Co-Ga-Zr-B alloy under a given
HDDR Process for the Production of High Performance...
Performance. . . 13

hydrogen pressure, there exists a certain temperature range in which the


residual Nd2z (Fe, Co, Ga)14 B phase is thermodynamically in equilibrium with
decomposition products of this phase, namely, ex-(Fe,Co), (Fe,Co)2B, and
o:-(Fe,Co), (Fe,Co)zB,
z , and claimed to have experimentally determined the temperature range
NdH 2
of such a four-phase equilibrium field. Fujita and Harris (1993) showed by
Thermopiezic Analysis (TPA), that increasing the Co content in
Ndzz (Fe1- xCox)
22 (Fel- 05 alloys, from x = 0 to O. 5, decreases the amount of
x Cox) 14 B I105
hydrogen absorbed during disproportionation.
Ikegami et al. (1996) also proposed a model of texture formation (shown
in Fig. 1.10), where if disproportionation is allowed to start in the four-phase
equilibrium field, some portion of the original Nd2 z (Fe, Co, Ga) 14 B remains
Ga)14
undisproportionated, maintaining the original crystallographic orientation. In
the recombination stage, the recombination proceeds mainly from the finely
dispersed Ndz2(Fe, Co, Ga) 14 B crystal Iites, resulting in a textured
microstructure with the original crystallographic orientation. Tomida et al.
(1997) observed nanometer size of residual Nd2z Fe14 Fel4 B particles during TEM
observations and suggested that they act as recombination centers leading to
the magnetic anisotropy formation.
] 1 - - -_____
rr------,

a

'B'

o
I

Figure 1.10
1. lOA Nd, (Fe, M)14B
A model of texture formation in Ndz(Fe, M) 14 B during the HDDR
process. The center is a model of a phase diagram on the hydrogen pressure-
temperature plane. The hatched area represents the NdH,NdH z + (Fe, M), B + (Fe, M) +
M) zB
Nd 14 B + H,
Nd,z (Fe, M) 14 Hz phase field (Ikegami et ai.,
aI., 1996).
1996)

Recently, Tomida et al. (1999) reported observing an "intermediate


hydrogenation phase" in the early stage of disproportionation of a
Nd13Fe679Co11
Nd 13 Fe679 COil 0oB70Ga1
B70 Gal 0OZrOl
Zro I alloy, which was identified as tetragonal Fe3 B
(t-Fe3 B). The t-Fe3 B phase possesses lattice coherency and an associated
"one-to-one" orientation relationship between the original Nd2 Fe14 B phase.
z Fe'4
14 Satoshi Sugimoto and David Book

Within the t-Fe3 B grain, 50 nm sized particles of Nd2Fe14B,


Fe14 B, which have a close
orientation to that of the original Nd2Fe14 B, exist alongside NdH 2 particles.
They suggested that the fine Nd2Fe 14B
Fe14 B particles act as recombination centers,
t-Fe3 B with Nd
and that the lattice coherency of the t-Fe3B 2Fe 14 B and its metastability
Nd2Fe14B
plays an important role in the mechanism of anisotropy during HDDR.
However, Nakamura et al. (1994) found that magnetic anisotropy could
be induced in a ternary Nd 126 12 6 FeSl
FeS144 B60 alloy, by raising the processing
temperature during the Desorption Recombination stage of the HDDR process.
This report contradicts the "four-phase region" model proposed by Uehara
et al. (1993) in order to try to explain the origin of anisotropy. As described
in Section 1.2.3,
1.2. 3, Nakamura et al. (1995) also suggested that the main effect
of additions is to alter the reaction kinetics of the disproportionation and
recombination reactions. The Zr or Nb addition makes the disproportionation
reaction sluggish and the Co or Ga addition decreases the temperature of the
recombination reaction. Sugimoto et a!. al. (1997) also reported the same effect
of additional elements by resistivity measurements during HDDR phenomena.
The report showed that the addition of Co and Ga changes both the hydrogen
pressure and temperature during disproportionation and recombination, and
therefore the rates of reaction.
Following up on the discovery that magnetic anisotropy could be induced
in a ternary alloy by raising the processing temperature during the
recombination reaction, a thermodynamic study was carried out by Nakamura
al. (1998, 1999a), which clarified the hydrogen pressure dependence of
et a!.
the recombination reaction temperature, for disproportionated Nd2Fe14 B
powder. Table 1. 1 shows the thermodynamic data with respect to H2 for
different Nd-Fe-B alloys, calculated using the starting temperature of
recombination. Using these data (solid circles), the relationship between the
hydrogen pressure and temperature during HDDR was clarified, as shown in
Fig. 1. 11. Also shown in this figure is data (open triangles) obtained by
Sugimoto et al.a!. (1997), who measured the recombination pressure by
monitoring the change of the electral resistivity during HDDR. It can be seen
that the recombination pressure of hydrogen is higher than the dissociation
pressure of NdH 2 , which means that the dissociation of NdH 2 is not the rate-
determining step in the recombination reaction. A lower boundary (denoted by
rhombi) was added to Fig 1. 11 using the data of the starting temperature of
the disproportionation reaction (heating from room temperature under
hydrogen), as reported by Book and Harris (1995). Therefore, the P- T plot
can be divided into three regions ( I - ill) as shown in Fig. 1. 11: ( I ) the
Nd2Fe14 B compound (or its hydride) is stable, (n) (II) the disproportionated
mixture is stable, and (ill) although the disproportionated mixture is more
stable, the Nd2Fe14 B compound exists because of unfavorable kinetics at these
low temperatures. In the typical HDDR process, Nd-Fe-B alloys go from
(ill)-( IIn)-(
)-( I).
HDDR Process for the Production of High Performance. . . 15

Iloor
1100 I B+H2
(Ill)Nd 2Fe I4 8
1000

900
1Nd2FeI4B+H2=NdH2+Fe+Fe2B+H2
Nd2FeI4B+H?=NdH2+Fe+Fe2B+H?
NdH2=Nd+H~
NdH 2=Nd+H 2 ",,"
/' /' /'
/">-//
/\//
-
f-!
:.J
h
f::'
i 800
1!
e~'" 700
:::l

~
E --~
--~
~
600 B +H 2
( I)Nd 2Fe I48
(J)Nd
500 L - L- -'---- ---'- ---.J

102
Hydrogen pressure, PH, (Pa)

Figure 1. 11 The temperature dependence of the recombination pressure of the Nd2 zFe14 8
compound. The dissociation pressure of NdH2 z and the starting temperatures of the
al. , 1998).
disproportionation reaction during heating, are also shown (Nakamura et aI.,

Table 1. 1
I Thermodynamic data with respect to H Hz2 for different Nd-Fe-8
Nd-Fe-B alloys, calculated
using the starting temperature of recombination. Data for Nd (Krost and Warf,
1966) is shown for comparison (Nakamura et al. , 1998).

--fJ.H
b.H --fJ.S
b.S "C P
b.G at 850 'C
- fJ.G PHHZ2 at 850 "C
'C
Composition
(kJ/mol of H
Hz)
2) (J/(K mol) of H
Hz)
2) (kJ/mol of H
Hz)
2) (kPa)

NdlZZ
Nd 12 .2Fea18
Fesl.8 8 60
so 186.0 146.2 21.80
2180 9.8
98
Nd,z.z
Nd Feao.asGalO
l22 Feso GalO 8 s6.0 193.2
1932 155.
155.1 19.00
1900 13.3
NdlZ.zFe7S0C05
NdI2 .2Fe76.0CoS a8s
sB6 0 1987
198.7 160.9 17.99 14.7
Nd 1Z .ZFeS43 C0 17 . 58 s 0
NdI2.2Fe64.3Co17.sB60 246.9
2469 217.9 2.166 80.2
Nd 2113
211.3 145.6 47.77 0.6
06

Nakamura et al. (1998, 1999a) also investigated the effect of additives


(Co, Ga) on the hydrogen pressure of the recombination reaction. Co and Ga
enhance the hydrogen pressure of the reaction as shown in Fig. 1. 12 and
Table 1. 1. The addition of Co and/or Ga increases the recombination
pressure, which was found, for instance, to be 80 kPa for 17.5 at% Co alloys
compared with 9.8 kPa for the ternary alloy, at 850 'C
"C .
Using the Pressure-Temperature curve shown in Fig. 1. 11, Nakamura
et al. (1998, 1999a, 1999b) and Sugimoto et al. (1999) proposed a type of
HDDR treatment for the production of anisotropic HDDR ternary powders. It
was a combination of heat treatments, at hydrogen pressures close to the
recombination pressure of the Ndz Fe14 B compound, in both the
disproportionation and recombination stages. The schematic illustration of the
newly proposed HDDR treatment is shown in Fig. 1.13,
1. 13, and is compared with
the conventional HDDR treatment (where c-HD is the hydrogen treatment for
the disproportionation stage and c-DR is the heat treatment in vacuum for the
16 Satoshi Sugimoto and David Book

1000 ....
.....
......
.................
,-.. 900
~ 900 ..........
l-J
h
i~.
::l
850 Ternary
.........
< ;
~ ~
1.0%Oa ~~'""'-
1.0%Ga; / '"'"
e
(;j
8.~ 800 ......... ~~~
/;/ \7.5%Co
17.5%Co
E ...........' ,..,
;

~
750
750 ~ ~ ~ ~ ~ ~ ~I NdHz+Fe+FezB+Hzl
700 5 6.;.... ~; 3 4 5 67
234567
Z 3 4 567
234567
0.01 0.1
0.\
Hydrogen pressure, PH,(Pa)

Figure 1. 12 The temperature dependence of the recombination pressure of


Nd12.2Fes1.8-xMxB6.0(Mx = C0 17S
Nd 12 . 2 Fes1.8-x M x 8 60 (M x =C0 17 .5 or Gal.O) (Nakamura et aI.,
al. , 1998).

recombination stage). Namely, it is a combination of v-HD (heating under


vacuum up to the treatment temperature, exposing to a mixture of hydrogen
and Ar for 1 h, and then exposing to O. 1 MPa of hydrogen for another in the
hydrogenation disproportionation stage) or I-HD (heating in a low H2 2 pressure

during the Hydrogenation Disproportionation stage), and s-DR (heating in Ar


or in hydrogen with a certain pressure for 10min, 10 min, at the start of the
Recombination) treatments. The high remanences of 1.3 - 1.4 T (B,/ (Brl J s > >
O. 9) after this new HDDR treatment, proved that additives are not necessary
to obtain anisotropic HDDR powder. More recently, other research groups
have reported similar HDDR treatments in Dy and Co or Ga and Nb added
samples (Morimoto et ai.,a!., 1999; Gutfleisch et ai., a!., 1999; Mishima et al. a!. ,
2000), with Mishima et al. a!. (2000) achieving a (BH ) max of 342 kJ/m 3
(43 MGOe). In addition, Hamada et al. (2003) reported a high performance
bonded magnet with (BH) max of 216 kJ/ kJI m3 .
Nakamura et al.
a!. (1998, 1999a) and Sugimoto et al. a!. (1999) observed
various microstructural changes during the new HDDR treatment. Figure 1. 14
shows optical and SEM microstructures of Nd12.2Fe818B60
Nd 12 . 2 FeslS 8 60 after c-HD a) and
(b))
(b and v-HD c) and (d)) (d treatments at 950 C . The optical micrograph in
Fig. 1. 14a shows that the c-HD sample has a relatively uniform microstructure
consisting of disproportionated mixture. In the SEM micrograph shown in Fig.
1. 14b, spherical NdH 2 ~m in diameter (denoted by "S") ,
2 grains of less than 1 IJm
embedded in an Fe matrix ("M") (" M") were observed, together with fine NdH 2 2

grains (" R"). This morphology is characteristic of the disproportionated


mixture (Nakamura et ai., al., 1996; Itakura, 1998). The optical micrograph
shown in Fig. 1. 14c reveals that the v-HD treatment results in quite a different
morphology. Coarse lamellae ("C") form a network, and fine lamellae with an
inter-lamellar spacing of about 250 nm (as shown by the close up of area "F"
HDDR Process for the Production of High Performance.
Performance...
.. 17

1100

1000
f..J
Y (c-DR)
h
h
900 ~~--==----="':;2~=
2
1:E
~
.'3 800
["0~.
~
~
"E 700

600
600 (III) Nd2Fe14B+H2
500
500 LL I . - -_ _- - ' - - '_ _- - ' ' - - -_ _- - '

1022 103 104 105 106


Hydrogen pressure, PH, (Pa)

,------------.\Treatment temp.:850-950 c
,---------\Treatment r---------,
'C } -------....,

Disproportionation stage Recombination stage

Temperature held
for tv until
pressur-I Pa
F,'2 =0.1 MPa .I'-DR treatment
v-HD treatment 0-90min

Ar
(QJn=500)
(Q,"=500)
or
Ic:::l
Const.
Followed by the
usual evacuation
treatment (c-DR).
Press.H 22

PH,=10-51 kPa

Figure 1.
I. 13 The hydrogen pressure-temperature curve of the recombination
Nd,2 Fe" B compound, the dissociation of NdH
reaction of the Nd NdH,2 and the starting
disproportionation reaction during heating. c-HD, v-HD and I-HD denote
conventional, vacuum and low-pressure disproportionation treatments,
respectively. While, c-DR and s-DR denote conventional and slow-rate
recombination treatments.

in Fig. 1. 14d), can be observed inside this network. Similar microstructures


have been reported by Gutfleisch et al. (1995b), and Gao et al. (1997). The
samples v-HD treated at 850
850Cc had morphologies similar to those treated at
950 c , however, the lamellar spacing was significantly smaller.
18
,8 Satoshi Sugimoto and David Book

(c) (d)

Figure 1. 14 Microstructures of Nd Nd'2 S 8 6 ..00 after disproportionation at 950 C. c-HD


,Fes1 8B
12 2Fe8'

treatment: (a)
Ca) optical microscope image, and (b) FE-SEM image. (Regions of NdH, NdH2 (S
and R) and Fe (M) are labelled.) v-HD treatment. (c) optical microscope image showing
fine and coarse lamellae regions and Cd) (d) FE-SEM image of a close up of the fine Lamellae
'998; Sugimoto et al. , 1999).
al. , 1998;
ragien. (Nakamura et al., '999).

Figure 1. 15 (Sugimoto et al.,


aI., 1999) is a schematic illustration of the
microstructural changes that occur during the disproportionation reaction for the
c-HD and v-HD treatments. In the c-HD sample, the disproportionation
reaction first takes place near the grain boundaries and hydrogen diffuses into
grains during heating in hydrogen, resulting in a uniform distribution of fine
lamellae. The disproportionation reaction is accompanied by a large volume
expansion (Nakamura et al., 1999c) and it is possible that the
disproportionated mixture prefers to form a lamellar structure in order to
reduce the strain due to the volume change. However, this results in a large
interfacial energy and the lamellae will then grow, with increasing
temperature, to form spherical grains of NdH 22 in order to reduce the interfacial
energy. The disproportionation reaction during v-HD treatment is thought to
begin at the grain boundaries. However because of the higher treatment
temperature, hydrogen diffusion into the grains and growth of the fine lamellae
occurs almost simultaneously, which leads to a more inhomogeneous
morphology than for c-HD treated samples. In addition, the coarse lamellae
form a network which could act as a diffusion path for hydrogen at the early
stage of the reaction.
Figure 1. 16 shows the remanence (B r ) and coercivity (jH c )) of Nd'2.2
Nd ,2 . 2
Fe818 8 60 alloys after HDDR treatment at 850 and 950 'c, versus s-DR
FeSI8
treatment time. The samples were disproportionated by either the (a) c-HD
treatment, or (b) v -HD treatment. Normal ized remanence (B r / J s ) , is
indicated by dashed lines for the values 0.5 and 1. O. The t = 0 min samples =
HDDR Process for the Production of High Performance. . . 19

IC-HD treatment I FHD!reatment


treatment I

O~
~~ (D~IOO~lm)
Nd 2 Fe l4 B (!:J Nd,Ie,,"
( { ) Nd,F'"B

n Heat

.
Heal in H
H?2
U
n
11
~.J
Heat in Vac. then
Expose to 11
E"pose H2


_600
6 00C
C 0

Dispropoliionation
Disproportionation
D.isPropoltionation reaction starts at the
'ii'';iii,.i,
:,'. ..'!..i..!."'.. .,... .. ... . . . Disproportionation
.!!
OJ> reaction starts at the boundary
";',.".". boundary

Disproportionation is completc DispropOliionation is complete


Disproportionation
Fine lamel.lae in I
ll'in n I Network lamellae,
NetlVork of coarse lamellae.
l small colonies. D ~vith
with fine lamellae inside

-~~800'C
-800 'C
~ ~ Spherical
Spherical NdH embedded in
NdH 12 embedded in Fe
Fe

~.
Figure 1. 15 Schematic illustration of microstructural changes during the
disproportionation reaction in a Nd-Fe-B alloy. for the c-HD and v-HD treatments
(Sugimoto et al
al ., 1999),

had no s-DR treatment, and so were only processed with the c-DR treatment.
850 c , the remanence
In the samples disproportionated by c-HD treatment at 850C,
O. 85 to 1,0
increases gradually from 0,85 1. 0 T with increasing ss-DR
-DR treatment time.
950 c , the remanence was higher up to an s-
With a treatment temperature of 950C,
DR treatment time of about 40 min,min. The coercivities were not very sensitive to
s-DR treatment time under these conditions. In contrast to the c-HD treated
samples, the v-HD samples exhibited greatly enhanced remanence due to the
s-DR treatment, as shown in Fig, Fig. 1.
1, 16b. A maximum value of 1.4 T
= 0.92) was obtained for the sample HDDR treated at 950
(B r / J 5 = c with an
950C
s-DR treatment of 20 min. Therefore, it may be said that the s-DR treatment
is an effective method of producing highly anisotropic Nd-Fe-B HDDR
powders, and that the conditions during the disproportionation stage also affect
the inducement of anisotropy.
anisotropy,
Sugimoto et al.
al (1999) also put forward a model to try to explain the
mechanism of anisotropy during HDDR. Figure 1. 17 shows a schematic
20 Satoshi Sugimoto and David Book

Nd l2 ZFeS18 8 6 0
1.6 Nd122Fe818B60 1.6 Nd122Fe818B60
~~H-Dt~;at~~~t-----~~/~~~---
------------------~--------- ------------------~---------
c-HD treatment B,lJs= 1 v-HD treatment B,/J = 1
1.4 1.4 s
Treattemp.,T("C )
Treattemp.,TCC
E 1.2 eT=950 'C
eT=950C E 1.2
~"
CQ"
<:Q
cr5" B,lJs=0.5
BPs=0.5 0T=850
T=850'CC
1.0

0.8 0.8 L~!!~~~:


L~!!~~~~ _

::~-----o-~~_
::~----o-,--=~-
I~0.5t
~ 1.5
..E Treattemp.,TCC
Treattemp.,T("C )
1.0
::;: eT=950'C
eT=950C

:~
2 oT=850'C
T=850 C
0.5~_.-~-_ _
~ 0.5L-_---~-
o 1 a 20 30 40 50 60 o 40 50 60
s-DR Treatment time, ((min)
t (min) Treatment time, t (min)
s-DR Treatmenttime,/(min)
(a) (b)

Figure 1. 16 Remanence and coercivity of Nd '22


12 . 2 FeS18
FeslS 8 60 alloys after HDDR treatment at
950 'c vs. s-DR treatment time, for samples that had been disproportionated by the
850 and 950C
(a) c-HD and (b) v-HD treatment. (Nakamura et aI., al. , 1998; Sugimoto et aI.,
al. , 1999).

illustration of the microstructure due to the difference in hydrogen pressure


during the recombination reaction. From the results described above, it may
be said that the hydrogen pressure during the recombination stage PH2 has as
associated free energy change I1G,
/::"G, which will affect the driving force of the
recombination reaction and the frequency of nucleation of the newly formed
Nd z Fe14 B grains. The value of I1G
Nd2Fe14B /::,.G will increase with decreasing P H2
H2 ,, below
the recombination pressure-temperature Iineline of the Nd2 z Fe14 B compound.
Assuming that the nucleation of Nd2
z Fel4
Fe14 B grains with the same orientation as
the original grains requires only a small free energy change, then the
enhancement in anisotropy after the s-DR treatment could be explained by the

v-DR(PH,>P2)
v-DR(PH,>P Z)
Oriented

@:;'i*1
~::'*
~ NdH2~
NdH2~
Fe
c-DR(PH,>P2
c-DR(PH,>P Z)
'-
4-
o Misoriented

~~-
<=
>.<:
u
'-' 0
<= .-
<:
'" 'iii
":::l <;;
:::l '"
<T-
0"- "
"'"
....
u.. <:
<=
u
'-'
:::l

In PH,
Figure 1. 17 Schematic illustration of the microstructure found due to the difference in
al. , 1999).
hydrogen pressure during the recombination reaction (Sugimoto et aI.,
HDDR Process for the Production of High Performance...
Performance. . . 21

(PHHZ2 > P 2 ) In contrast, a lower hydrogen pressure


higher hydrogen pressure (P
(P H2
HZ < P2 ) would result in a larger free energy change, and therefore the
formation of a greater number of disoriented nuclei. Sugimoto et al. (1999)
concluded that the combination of v-HD (or I-HD) and s-DR treatments is an
effective way of obtaining highly anisotropic Nd-Fe-B HDDR powders.
a!. (2002) reported that the main tenance of
Very recently, Sugimoto et al.
crystal orientation relationship, between the disproportionated, mixture and
the Nd2Fel4 B phase during HDDR is important for the inducement of anistropy.
al. (2002,2003) proposed the Fe2 B phase in the disproportionated
Gutfleish et a!.
mixture as the anisotropy-mediating phase.

1. 4 HDDR Phenomena in Other Rare-Earth Magnetic


Materials

In 1990, Coey and Sun (1990) discovered that the magnetic properties of the
Sm2 Fel7 phase are dramatically improved by the introduction of nitrogen. The
nitrogen induced Sm2 Fell Fel7 compound was prepared by heating Sm2 Fell at
450 - 500 c C under a nitrogen or ammonia gas. The crystal lattice expansion of
the Sm2 Fell
Fel7 by mdre than 6 % to accommodate three nitrogen atoms in
interstitial sites, is accompanied by a dramatic increase in Curie temperature
T c' from 125 to 479 C. The saturation magnetization of Sm2 Fell Fel7 N3 (J s =
1. 54 T) is comparable to that of Nd2Fel4 B and the uniaxial magnetic
anisotropy (/-IoH
(J.l o H A =26 T) is three times as strong as that of Nd2Fel4B (Fujii
and Sun, 1995).
Nakamura et al. a!. (1991, 1992) and Sugimoto et al. (1992) were the first
to demonstrate that the HDDR process could also be appl ied to the Sm2 Fell
applied Fel7 Nx
compound. The hydrogen absorption and desorption characteristics of the
Fel7 compound were investigated, as shown in Fig. 1. 18. There are two
Sm2 Fell
absorption peaks (one around 250 - 350 c -- 600 C),
C and the other at '" 'C), and
two desorption peaks (one centered at 350 - 550 c C and the other at 1050-
1100 c ). X-ray diffraction showed that the second absorption peak at around
C ).
600 c
C corresponds to the disproportionation of Sm2 Fel7 into a-Fe ex-Fe and SmH2
phase, and that the second desorption peak at around 1050 - 1100 c C
corresponds to the recombination into the Sm2 Fel7 compound.
The HDDR processing of Sm2 Fell' Fe17' involves heating the alloy in hydrogen
to about 800 c, C, followed by a vacuum heat treatment to recombine the
Fel7 compound. Then after nitrogenation at 500 c
Sm2 Fell C for 4 h, the HDDR
processed powder was found to exhibit a coercivity of about J.lo =
/-10 He = 0.87 T.
22 Satoshi Sugimoto and David Book

o Sm2FeI7(HJ
Sm2 Fe l7 Hx
H treatment(T,OClh)~ a-Fe
vSmH 2

- --
Heating rate:400 'C/h
C/h
Particle size:<63 !lm
J.1m

(b)
(c)
(~)-A
(~)~ ~
!\ ./
./
v
V il''if
(d) (d) 1100'C
-.J\ l'<?
300 600 900 1200 40 50 60 70
Temp. eC)
TempCC) 2en
28 (')
(a) (b)

Figure 1. 18 (a) The characteristics of hydrogen absorption and desorption for Sm2 8m2 Fell
Fe17
powders during heating at a rate of 400 'C
C /h in hydrogen. (b) X-ray diffraction patterns of
8m2 Fell
Sm2 Fel7 powders after heat treatment at 1h in hydrogen at 270, 450, 775 and 1100 'C C
(Sugimoto et ai., 1992).

This showed that the HDDR process is a useful way of producing high-
coercivity. isotropic Sm2 Fel7 Nx powders. Christodoulou and Takeshita (1993)
coercivity,
also reported a high coercivity of O. 82 T. T, while Dempsey et al. (1996)
achieved coercivities up to 3. 3 T by combining HDDR with a milling
procedure.
Zhou et al. (1992) showed the effect of additional elements such as Cr or
Ga on the enhancement of coercivity. coercivity, with Sm2(Fe095Cro05)ll
Sm2(Fe095Cro05)l7 and
Sm2 (Feo 983 GaO
(Feo. 983 Gao 017
017)) 17 alloys exhibiting coercivities of 2.0 and 2.5 T.
T, respectively.
respectively,
after HDDR treatment and following nitriding. Nakamura et al. (Nakamura et al. ,
1995) reported the influence of additional elements on the HDDR conditions. conditions, in
which Zr or Nb addition makes the disproportionation reaction sluggish and Co or Ga
addition decreases the temperature of the recombination reaction under O. 1 MPa of
hydrogen. This influence of additional elements is similar to that seen for the HDDR
process in Nd-Fe-B, as described in Section 1. 2. 3.
Okada et al. (1992) investigated HDDR phenomena in the Sm2+5Fe17 Sm2+5Fell non-
stoichiometric alloys and magnetic properties of HDDR treated alloys after
nitriding. The nitrided Sm32 Fel7 powders exhibited a coercivity of 1.6 1.6 T.
T, and
microstructural changes were observed during the HDDR of the Sm2 Fell Fel7
compound. It was found that the microstructure of the disproportionated
Fel7 alloy consisted of spherical or rod-like SmH 2 phases several
Sm2 Fell
nanometres in size.size, within an ex-Fe matrix phase. Similar observations were
made by Clarke et al. (1996) using high resolution SEM. Okada et al. (1992)
HDDR Process for the Production of High Performance. . . 23

described the crystallographic relationship between the disproportionated


phases by <
(111 )a-Fe II <
111>a-Fell (110)smH
110)SmH2 ' Recombination was thought to initiate after
hydrogen was desorbed from SmH 2 phases, with a-Fe then diffusing into SmH 2
particles to form Sm2+5
Sm2+S Fel7
Fe17 (Okada et al., ai., 1995). However, recently, Liu
et al. (2000) reported that all the SmH 2 fibers have the same orientation
and grow with a definite orientation relationship to the a-Fe matrix i. e. ,
{ 100 }SmH 2 II {100}a_Fe
{100}SmH {1 OO} a-Fe and < 00 1) SmH 2 II <
(00l>SmH (OOl>a-Fe'
00 1) a-Fe'
1999,2000)
Recently, Tobise et al. (1998, 1999, 2000) obtained magnetic materials
with high magnetic coercivity in the Sm-Fe-B-Ti alloys prepared by HDDR,
followed by nitriding. The addition of a small amount of Ti and B in Sm-Fe-N
were effective in eliminating
el iminating a-Fe, and obtaining isotropic Sm94 Fe869
FeS69 BTi u Ns6
powders with the following magnetic properties: remanence B r = = O. 76 T,
~oHc=l.l
/-10 He = (BH)max=88
1.1 T and (BH)max = 88 kJ/m 3 CTobise et ai.,
al. , 1998). It has been also
found that the magnetic characteristics of Sm-Fe-B-Ti alloys do not decrease
even after being exposed to high temperatures. Temperature coefficients of B r
~oHc
and /-10 He of the Sm-Fe-Ti-B-N bonded magnet were - 0.06 % I'C
-0.06%/"C
and - O. 43 % Fc,
/"C, respectively in the temperature range 25 to 100C. 100'C. As
shown in Fig. 1. 19, the irreversible loss for the Sm-Fe-Ti-B-N bonded magnet at
100C
100 'c is - 2. 1%
1% and - 2 2.. 5 %
% after exposure for 2 and 300 h, respectively. This
irreversible loss is lower than that for Sm-Fe-N CTobise et ai., 1999).
Time(h)
10 100 1000

Exposure temperature: I00 'C

:E -4
:E-4
.~
OJ
<I)
>
~~ -6 SmFeN(11 kOe)
SmFeN(16kOe)
SmFeTiBN(1IIJ kOe)
o SmFeTiBN(

-8 I 10 100 1000 10000


Time (ks)
Figure 1. 19 The change of irreversible loss of the Sm-Fe-Ti-B-N and Sm-Fe-N
bonded magnets as a function of exposure time (T obise et al. , 1999).
(Tobise

1. 4. 2 ThMn12-type Compound

After the discovery of the Nd2Fe14 B compound, much attention was paid to the
R(Fe, M)12 compounds (R = = rare earth and M == Ti, V, Mo, Si, etc.) as
potential permanent magnetic materials. As the RFe12 compounds do not exist,
24 Satoshi Sugimoto and David Book

additions of a third element are needed to stabilize the ThMnl2-type ThMn12 -type body-
centered tetragonal structure. Of particular note are the Sm ( Fe, M) 12
compounds, which have a high uniaxial magnetic anisotropy and a relatively
high Curie temperature. Magnetic hardening of this type compound was
investigated mainly by melt-spinning (Ohashi et al., aI., 1987; Singleton et al. al.,,
1989; Okada et al.. ' 1989; Okada et al., 1990; Ding 0 ing and Rosenberg, 1990)
and mechanical alloying (Schultz and Wecker, 1988). However, Okada et al.
(1992) and Tatsuki et al. (1993) found that the HDDR process could be
applied to the Sm(Fe,M)12 compound, and that isotropic SmFe,o SmFelO TiV powders
O. 6 T could be obtained.
with a coercivity of 0.6
After the discovery of the excellent magnetic properties of Sm2 Fel7 Fe17 Nx ,
interstitial modification studies were extended to the ThMnlrtype compound.
In this compound nitrogen atoms are located in the octahedral sites, resulting
in an increase in Curie temperature of about 200 'C. The Nd-based ThMnlr ThMn12-
type compounds exhibited a particularly strong uniaxial anisotropy and seemed
to have the most potential for use as permanent magnets (Yang et al. , 1991;
Fujii
Fuj ii and Sun, 1995). Tatsuki
Tatsuk i et al. (1993) observed the occurrence of HDDR
phenomena in NdFelOMM'
NdFelO MM' (M = Ti, V, Mo, M' =V,
(M=Ti, = V, Mo) alloys, however the
NdFell Ti compounds recombined into a TbCu7 -type structure after HDDR.
They also obtained a high coercivity of 0.58 T with Nd l 3Fe,o 3FelO VMoN x .' Sugimoto
et al. (1994) investigated the influence of the third elements, which stabilize
the ThMnl2
ThMn12 phase, on hydrogen absorption and desorption characteristics. Mo
makes the disproportionation reaction sluggish and shifts disproportionation to
a higher temperature, whereas Co addition has the effect of lowering the
temperature of recombination. It was also shown that Co addition increases
the coercivity, with isotropic Nd 16 (Feu(Feo9 Coo
COO. 1) 10 V2 Nx powders exhibiting a
1 ) 10

O. 83 T.
coercivity of 0.83

1. 4. 3 Nd3 (Fe, Ti)29-Type Compound

ColiocoU
Collocott et al. (1992) reported a new intermetall
intermetallic
ic compound which has been
shown to have a monoclinic structure and a composition of R3(Fe, M)29(where
R== rare earth, M = = transition metal) (Cadogan et al., 1994; L Lii et al. ,
1994). The composition lies between the 1-12 and 2-17 phases, and can be
viewed as an alternating stack of 1-12 and 2-17 units (Li et al.,
aI., 1994). As
with R2Fell
Fe17 compounds, nitriding greatly improves the magnetic properties,
and in particular, Sm3 (Fe, M) 29 Nx is a very promising alloy for permanent
magnets (Nasunjilegal et al. , 1995; Suzuki et aI.,
al. , 1994).
Book et al. (1996a, 1996b) studied HDDR phenomena in the
Sm3 (Fe, V)29 compound. XRD measurements showed that the samples did not
recombine into Sm3 (Fe, V )29' but rather, recombined into a mixture of a
V)29'
HDDR Process for the Production of High Performance. . . 25

TbCu7-type phase and cx-Fe(V).


o:-Fe(V). After HDDR treatment at 800 C and nitriding
at 560 C for 4 h, the materials exhibited coercivities up to 0.73 T. They also
investigated the effect of Co addition on the grain size of the recombined grains
(Book et al. 1998, 1999).

1. 5
1.5 Summary

The HDDR process has proved to be an effective way of producing coercive


Nd-Fe-B powder, suitable for anisotropic bonded magnets, and has also been
applied to certain Sm-Fe compounds, such as Sm2 Fel7Fe17 Nx and SmFelO TiV, to
obtain coercive, isotropic powder.
Anisotropic Nd-Fe-B powders can be produced using just the ternary alloy-
additives such as Zr and Ga are not necessary - by carefully controlling the
rates of reaction at the beginning of disproportionation and recombination
(e. g. , by using the I-HD and s-DR treatments). However, additives can be
useful for changing the HDDR conditions (i. e., the pressure and/or
temperature at which disproportionation and recombination occur), and for
preventing excessive grain growth during recombination which can cause a
decrease in coercivity.
The mechanism of the inducement of anisotropy during HDDR is still
unclear, however, it can be said that the state of the disproportionated
mixture is the key: a fine lamellar structure of Nd hydride, as opposed to
spherical particles, will result in highly orientated Nd2 Fel4
Fe14 B grains after
recombination (Sugimoto et al., 1999). Another point to consider is that
HDDR-induced anisotropy has only been observed in Nd2Fe Fel4
14 B-based alloys,
and not in any other magnet alloys such as Sm2 Fel7 Nx . This suggests that the
presence of iron-boride phases (Fe2 B or t-Fe3 B) in the disproportionated
mixture (Tomida et ai., 1999; Gutfeish et ai., 2002, 2003), and/or
differences in the crystal structure of the magnetic phase (Sugimoto et al. ,
2002), could be important factors in the promotion of anisotropy during
recombination.

References
Book, D. and I. R. Harris. IEEE Trans. Magn. 28: 2145 (1992)
Book, D. and I. R. Harris. J. Alloys Comp. 221: 187 (1995a)
Book, D. , I. R. Harris, A. Manaf, I. Ahmad and H. A. Davies. J. Alloys
Compo 221: 180 (1995)
Book, D., H. Nakamura, S. Sugimoto, T. Kagotani, M. Okada and M.
Homma, Mater. Trans. JIM 37: 1228 (1996a)
Book, D., H. Nakamura, S. Sugimoto, T. Kagotani, M. Okada and M.
26 Satoshi Sugimoto and David Book

Homma. In: Proc. of the 14 th Int. Workshop on Rare-Earth Magnets and


Their Applications, Sao Paulo, Brazil (World Scientific, Singapore,
1996b), p. 138
Book, D. , K. Kato, H. Nakamura, S. Sugimoto, T. Kagotani, M. Okada
and M. Homma. In: L. Schultz, K. -H. - H. Muller eds. Proc. ofthe
of the 15th Int.
Workshop on Rare Earth Magnets and Their Applications. Dresden,
Germany (Werkstoff-Informationsgesellschaft, Frankfurt, 1998), p. 543
Book, D. , K. Kato, H. Nakamura, S. Sugimoto, T. Kagotani, M. Okada
and M. Homma. J. Magn. Soc. Japan 23: 314 (1999)
Burkhardt, C. , R. S. Mottram, F. Dime,
Dimc, S. Kobe and I. R. Harris. In: L.
Schultz, K. -H. Muller eds. Proc. of the 15th Int. Workshop on Rare Earth
Magnets and Their Applications. Dresden, Germany (Werkstoff-
Informationsgesellschaft, Frankfurt, 1998), p. 953
Buschow, K. H. J. IEEE Trans. Magn. 30: 565 (1994)
Buschow, K. H. J. In: K. H. J. Buschow ed. Handbook of Magnetic
Materials Vol. 10, Chapter 4. Elsevier Science B. V. , p. 463 ( 1997)
Cadogan, J. M., H. -So Li, A. Margarian, J. B. Dunlop, D. H. Ryan, S.
J. Collocott and R. L. Davis. J. Appl. Phys. 76: 6138 (1994)
Cadogan, J. M. and J. M. D. Coey. Appr. Appl. Phys. Lett. 48: 442, (1986)
Capehart, T W., R. K. Mishra and F. E. Pinkerton. J. Appl. Phys. 73:
6476 (1993)
Christodoulou, C. N. and T. Takeshita. J. Alloys Compo 196: 155 (1993)
Clarke, J. C. , O. Gutfleisch, S. A. Sinan and I. R. Harris. J. Alloys Comp.Compo
232: L 12 (1996)
L12
Coey, J. M. D. and H. Sun.SUh. J. Magn. Mater. 87: L251 (1990)
CollocoU,
Collocott, S. J.,
J. , R. K. Day, J. B. Dunlop and R. L. Davis. In: Proc. of
the 7th Int. Symposium on Magnetic Anisotropy and Coercivity in Rare-
Earth Transition Metal Alloys. Canberra, Australia,
Austral ia, p. 437 (1992)
p.437
Croat, J. J. , J. F. Herbst, R. W. Lee and F. E. Pinkerton. J. Appl. Phys.
55: 2078 (1984a)
Croat, J. J., J. F. Herbst, R. W. Lee and F. E. Pinkerton. Appl. Phys.
Lett. 44: 148 (1984b)
Dempsey, N. M., P. A. P. Wendhausen, B. Gebel, K. -H. Muller and J.
M. D. Coey. In: Proc. of the 9th Int. Symp. on Magnetic Anisotropy and
Coercivity in Rare-Earth Transition Metal Alloys. Sao Paulo, Brazil.
(World Scientific, Singapore, 1996)
1996),,p.
p. 349
Ding, J. and M. Rosenberg. J. Less-Common Met. 166: 313 (1990)
Fastenau, R. H. J. and E. J. van Loenen. J. Magn. Magn. Mater. 157/158:
1 (1996)
Fujii, H., and H. Sun. In: K. H. J. Buschow ed. Handbook of Magnetic
Materials, Vol.
Vo1.9,9, Chapter 3, Elsevier Science B. V. V.,, p. 83 ( 1995)
p.83(1995)
Fujita, A. and I. R. Harris. IEEE Trans. Magn. 29: 2803 (1993)
Gao, J. X. Song and X. Wang. J. Alloys Compo 248: 176 (1997> (1997)
0.,, N. Martinez, I. R. Harris, M. Matzinger and J. Fidler. In:
Gutfleisch, O.
HDDR Process for the Production of High Performance. ..
.. 27

Proc. of the 8th Int. Symp. on Magnetic Anisotropy and Coercivity in


243 (1994)
Rare-Earth Transition Metal Alloys. Birmingham, UK, p. 243( 1994)
Gutfleisch, O. and I. R. Harris. J. Mater. Sci. 30: 1397 (1995a)
Gutfleisch, O. , M. Matzinger, J. Fidler and I. R. Harris. J. Magn. Magn.
Mater. 147: 320 (1995b)
Gutfleisch, O. and I. R. Harris. J. Phys. D:0: Appl. Phys. 29: 2255 (1996)
Gutfleisch, O. and I. R. Harris. In: L. Schultz, K. - H. MOiler
Muller eds. Proc.
of the 15th Int. Workshop on Rare-Earth Magnets and Their Applications,
Vol. 1. Dresden, Germany, p. 487 (1998)
Gutfleisch, 0., B. Gebel, M. Kubis, K. H. Muller and L. Schultz. IEEE
Trans. Magn. 35: 3250 (1999)
Gutfleisch, O. , G. Drazic,
Orazic, C. Mishima and Y. Hankura. IEEE Trans. Magn.
38: 2958 (2002)
Gutfleish, 0., K. Khlopkov, A. Teresiak, K. H. Muller, G. Drazic, Orazic, C.
Mishima and Y. Honkura. IEEE Trans. Magn. 29: 5926 (2003)
Hamada, N., N. , C. Mishima, H. Mitarai, and Y. Honkura. IEEE Trans Magn.
39: 2953 (2003)
Harris, I. R., C. Noble and T. Bailey. J. Less-Common Met. 106: L 1
(1985)
Harris, I. R.,
R. , P. J. McGuiness, O.
D. G. R. Jones and J. S. Abell. Phys. Scr.
T 19: 435 (1987)
Harris, I. R. J. Less-Common Metals 131: 245 (1987)
Harris, I. R. , and P. J. McGuiness. In: Proc. of the 11 th Int. Workshop on
Rare-Earth Magnets and Their Applications. Pittsburgh, p. 29 ((1990)
1990)
Harris, I. R. In: Proc. of the 12th Int. Workshop on Rare-Earth Magnets and
Their Applications. Canberra, p. 347( 1992)
Harris, I. R. , O. M. Ragg, G. Keegan and H. Nagel. In: Proc. of the 3rd
Int. Symp. on Physics of Magnetic Materials. Seoul, Korea, p. 638 ((1995)
1995)
Harris, I. R. In: J. M.D.
M. D. Coey ed. Rare-Earth Iron Permanent Magnets.
Oxford University Press, New York, (1996) Chapter 7
Herbst, J. F. Reviews of Modern Physics 63: 819 (1991)
(1991>
Ikegami, T., H. Tomizawa and S. Hirosawa. In: Proc. of the 9th Int.
Symp. on Magnetic Anisotropy and Coercivity in Rare-Earth Transition
Metal Alloys. Sao Paulo, Brazil (World Scientific, Singapore, 1996),
p.288
Isnard, 0., W. B. Yelon, S. Miraglia, and D. O. Fruchart. J. Appl. Phys. 78:
1892 (1995)
Itakura, M. , N. Kuwano, K. Yamguchi, T. Yoneki, K. Oki, R. Nakayama,
N. Komada and T. Takeshita. Mater. Trans. JIM 39: 95 (1998)
Kaneko, Y. In: H. Kaneko, M. Homma and M. Okada eds. Proc. of the 16th
Int. Workshop on Rare-Earth Magnets and Their Applications, Sendai,
Japan (The Japan Institute of Metals, Sendai, 2000), p. 83
Kaneko, Y. In: P. de Rango ed. Proc. of the 18th Int. Workshop on High
Performance Magnets and Their Applications. Annecy, France (2004).
28 Satoshi Sugimoto and David Book

p. 40
Kobayashi, K. In: Proc. of the 13th Int. Workshop on Rare-Earth Magnets
and Their Applications, Birmingham, UK, p. 717 (1994)
Krost, W. L. and J. C. Wart.Warf. Inorg. Chem. 5: 1719 (1966)
I' Heritier P., P. Chaudouet, R. Madar, A. Rouault, J. Senateur and R.
I'Heritier
Fruchart, C. R. Acad. Sc. (Paris) 299 ( II) (13): 349 (1984) (In French)
Li, H. S.,S. , J. M. Cadogan, R. L. Davis, A. Margarian and J. B. Dunlop.
Sol id State Commun. 90: 487 (1994)
Liu, Z.,
Z. , T. Ohsuna, K. Hiraga and M. Tobise. In: H. Kaneko, M. Homma
and M. Okada. Proc. of the 16th Int. Workshop on Rare-Earth Magnets
and Their Applications. Sendai, Japan (The Japan Institute of Metals,
Sendai, 2000), p.767
p. 767
Lisert, S., D. Fruchart, P. de Rango and J. L. Soubeyroux. J. Alloys
Compo 253-254: 140 (1997)
Luo, Y. Magnews: International Newsletter of the UK Magnetics Society,
Summer, p. 13( 1999)
McGuiness, P. J., X. J. Zhang, X. Y. Yin and I. R. Harris. J. Less
Common Met. 158: 359 (1990a)
McGuiness, P. J., X. J. Zhang, H. Forsyth and I. R. Harris. J. Less-
Common Met. 162: 379 (1990b)
Meisner, G. P. and V. Panchanathan. J. Appl. Phys. 74: 3514 (1993)
Mishima, C. , H. Hamada, H. Mitarai and Y. Honkura. In: H. Kaneko, M.
Homma and M. Okada. Proc. of the 16th Int. Workshop on Rare-Earth
Magnets and Their Applications. Sendai, Japan (The Japan Institute of
Metals, Sendai, 2000), p. 873
Mishra, R. K., G. Thomas, T. Yoneyama, A. Fukuno and T. Ojima. J.
Appl. Phys. 52: 2518 (1981)
Mooij, B. D. and K. H. J. Buschow. Philips J. Res. 42: 246 (1987)
Morimoto, K. , R. Nakayama, K. Mori, K. Igarashi and Y. Ishii. IEEE Trans.
Magn. 35: 3253 (1999)
Nakamura, H., K. Kurihara, T. Tatsuki, S. Sugimoto, M. Okada and M.
Homma. J. Magn. Soc. Jpn. 16: 163 (1991) (In Japanese)
Nakamura, H., S. Sugimoto, M. Okada and M. Homma. Mater. Chem.
Phys. 32: 280 (1992)
Nakamura, H., R. SuefujSuefuji,
i, S. Sugimoto, M. Okada and M. Homma. J.
Appl. Phys. 76: 6828 (1994)
Nakamura, H., S. Sugimoto, T. Tanaka, M. Okada and M. Homma. J.
Alloys Camp.
Compo 222: 136 (1995)
Nakamura, H. , R. Suefuji, D. Book, T. Kagotani, S. Sugimoto, M. Okada
and M. Homma. Mater. Trans. JIM 39: 95 (1996)
Nakamura, H. , K. Kato, D. Book, S. Sugimoto, M. Okada and M. Homma.
In: L. Schultz, K. - H. MOiler
Muller eds. Proc. of the 15th Int. Workshop on
Rare-Earth Magnets and Their Applications. Dresden, Germany (Werkstoff-
Informationsgesellschaft, Frankfurt, 1998), p. 507
HDDR Process for the Production of High Performance. . . 29

Nakamura, H. , K. Kato, D. Book, S. Sugimoto, M. Okada and M. Homma.


J. Magn. Soc. Japan 23: 300 (1999a)
Nakamura, H. , K. Kato, D. Book, S. Sugimoto, M. Okada and M. Homma.
IEEE Trans. Magn. 35: 3274 (1999b)
Nakamura, H. , K. Kato, D. Book, S. Sugimoto, M. Okada and M. Homma.
T. lEE Japan 119-A (6): 808 (1999c) (In Japanese)
Nakayama, R. , T. Takeshita, M. Itakura, N. Kuwano and K. Oki. J. Appl.
Phys. 70: 3770 (1991)
Nasunjilegal, B., F. M. Yang, N. Tang, W. D. Qin, J. L. Wang J. J. Zhu,
H. Q. Guo, B. P. Hu, Y. Z. Wang and H. S. Li. J. Alloys. Compo 222:
57 (1995)
Nesbit, E. A., R. H. Willens, R. C. Sherwood, F. Buehler and J. H.
Wernick. Appl. Phys. Lett. 12: 361 (1968)
Oesterreicher, H. In: Proc of
at the 4th Int. Symp. on Magnetic Anisotropy and
Coercivity in Rare-Earth Transition Metal Alloys. Dayton, USA p. 507
(1985
(1985))
Oesterreicher, K. and H. Oesterreicher. Phys. Status Sol idi A 85: K61
Solidi
(1984)
Ohashi, K., T. Yokoyama, R. Osugi and Y. Tawara. IEEE Trans. Magn.
MAG-23: 3101 (1987)
Okada, M., K. Yamagishi and M. Homma. Mater. Trans. JIM 30: 374
(1989)
Okada, M. , A. Kojima, K. Yamagishi and M. Homma. IEEE Trans. Magn.
26: 1376 (1990)
Okada, M. , S. Sugimoto and M. Homma. In Ferrites: Proc. 6th Int. Conf.Cant.
On Ferrites (/CF6). Tokyo and Kyoto, Japan (The Japan Society of
Powder and Powder Metallurgy, Tokyo), (1992), p. 1087
Okada, M. , K. Saito, H. Nakamura, S. Sugimoto and M. Homma
Homma. J. Alloys
Compo 231: 60 (1995)
Pollard, R. J. and H. Oesterreicher. IEEE Trans. Magn. MAG-22: 735
(1986)
Sagawa, M., S. Fujimura, M. Togawa and Y. Matsuura. J. Appl. Phys. 55:
2083 (1984)
Scholz, U. D., W. E. Kronert and H. Nagel. In: Proc of at the 9th Int.
Workshop on Rare-Earth Magnets and Their Applications. Bad Soden,
Germany, p. 26( 1987)
Schultz, L. and J. Wecker. J. Appl. Phys. 64: 5711 (1988).
Singleton, E. W., J. Stzeszewski and G. C. Hadjipanayis. Appl. Phys.
Lett. 54: 1934 (1989)
Skomsk i, R. In: J. M. D. Coey ed. Rare-Earth Iron Permanent Magnets.
Skomski,
Clarendon Press, Oxford, Chapter 4 4(( 1996)
Strnat, K. J. , G. Hoffer, J. Olsen, W. Ostertag and J. J. Becker. J. Appl.
Phys. 38: 1001 (1967)
Strnat, K. J. In: E. P. Wohlfarth and K . H. J. Buschow ed. Ferromagnetic
30 Satoshi Sugimoto and David Book

Materials, Vol 4 (Elsevier Science Publishers B. V. , North-Holland, 1988),


p.131
Strnat, K. J. and R. M. W. Strnat. J. Magn. Magn. Mater. 100: 38 (1991)
Sugimoto, S. , H. Nakamura, M. Okada, and M. Homma. In: Proc. of the
12th Int. Workshop on Rare-Earth Magnets and Their Applications.
Canberra, Australia, p. 372( 1992)
Sugimoto, S. , T. Tatsuki, H. Nakamura, M. Okada and M. Homma. In: M.
Homma et al. ed. Advanced Materials' 93, 1/ B: Magnetic, Fu/lerene, Fullerene,
Dielectric, Ferroelectric, Diamond and Related Materials. Trans. Mat.
Jpn. , Vol. 14B, Elsevier Science B. V.
Res. Soc. Jpn., V.,, p. 1041( 1994)
1041(1994)
Sugimoto, S., O. Gutfleisch and I. R. Harris. J. Alloys Comp. 260: 284
( 1997)
Sugimoto, S. , S. Ohga, K. Inomata, K. Suzuki, T. Konno, and K. Hiraga,
IEEE Trans. Magn. 38: 2961 (2002)
Sugimoto, S. , H. Nakamura, K. Kato, D. Book, T. Kagotani, M. Okada
and M. Homma. J. Alloys. Compo 293-295: 862 (1999)
Suzuki, S., S. Suzuki and M. Kawasaki. J. Appl. Phys. 76: 6708 (1994)
Takeshita, T. and R. Nakayama. In:. Proc. of the 10th Int. Workshop on
Rare-Earth Magnets and Their Applications. Kyoto, Japan, p. 551 (1989)( 1989)
Takeshita, T. and R. Nakayama. In: Proc. of the 11 th Int. Workshop on
Rare-Earth Magnets and Their Applications. Pittsburgh, USA, p. 49( 49 ( 1990)
Takeshita, T. , and R. Nakayama. In: Proc. of the 12th Int. Workshop on
Rare-Earth Magnets and Their Applications. Canberra, Australia,
Austral ia, p. 670
(1992)
Takeshita, T., and R. Nakayama. J. Mag. Soc. Jpn. 17: 25 (1993) (In
Japanese)
Tatsuki, T., H. Nakamura, S. Sugimoto, M. Okada and M. Homma. J.
Magn. Soc. Jpn. 17: 165 (1993) (In Japanese)
Tawara, Y. and H. Senna.
Senno. Jpn. J. Appl. Phys. 7: 966 (1968)
Tobise, M.M.,, M. Shindoh, H. Okajima, K. Iwasaki, M. Tokunaga, Z. Liu and
K. Hiraga. In: L. Schultz, K. - H. MOiler
Muller eds. Proc. of the 15th Int.
Workshop on Rare Earth Magnets and Their Applications. Dresden,
Germany (Werkstoff-Informationsgesellschaft, Frankfurt, 1998), p. 517
Tobise, M., M. Shindoh, H. Okajima, K. Iwasaki, M. Tokunaga, Z. Liu and
K. Hiraga. IEEE Trans. Magn. 35: 3259 (1999)
Tobise, M., Z. Liu, M. Shindoh, S. Tanigawa, and K. Hiraga. In: H.
Kaneko, M. Homma and M. Okada eds. Proc. of the 16th Int. Workshop on
Rare-Earth Magnets and Their Applications. Sendai, Japan (The Japan
Institute of Metals, Sendai, 2000), p. 793
Tomida, T.,
T. , P. Choi, Y. Maehara, M. Uehara, H. Tomizawa, S. Hirosawa.
Compo 242: 129 (1996)
J. Alloys Camp.
Tomida, T.,
T. , N. Sana,
Sano, and M. Uehara. J. Appl. Phys. 81: 7170 (1997)
Tomida, T. , N. Sana,
Sano, K. Hanafusa, H. Tomizawa and S. Hirosawa. Acta
Mater. 47: 875 (1999)
HDDR Process for the Production of High Performance. . . 31

Uehara, M. , H. Tomizawa, S. Hirosawa, T. Tomida and Y. Maehara. IEEE.


Trans Magn. 30: 2770 (1993)
Wiesinger, G. , G. Hilscher and R. Gr6ssinger.
Grossinger. J. Less-Common Metals 131:
409 (1987)
Xiao, Y., J. Liu, B. Qiu, and M. Liu. In: Proc. of the 12th Int. Workshop
on Rare-Earth Magnets and Their Applications. Canberra, Australia,
Australia , p. 258
(1992)
Yang, Y. C. , X. D. Zhang, L. S. Kong, Q. Pan and S. L. Ge. Appl. Phys.
Lett. 58: 2042 (1991)
Yu;
Yu" Z. , J. Zhang and Y. Zheng Hyperfine Interactions 68: 295, (1991)
Zhang, X. J., P. J. McGuiness, and I. R. Harris. J. Appl. Phys. 69: 5838
( 1991)
Zhou, S. Z. , J. Yang, M. C. Zhang, D. Q. Ma, F. B. Li and R. Wang. In:
Proc. of the 12th Int. Workshop on Rare-Earth Magnets and Their
Applications. Canberra, Australia, p. 44( 1992)

The authors would like to thank Professors Motofumi Homma (Professor Emeritus of Tohoku
University), Masuo Okada CTohoku University), Koichiro Inomata CTohoku University), Rex
Harris (University of Birmingham), Dr. Hajime Nakamura (formerly at Tohoku University,
now at Shin-etsu Chemical Co. Ltd) and Dr. Oliver Gulfleisch
Gutfleisch (IFW Dresden) for many
stimulating and helpful discussions. Some of the work featured in this chapter was supported
in part by a Grant-in-Aid for Scientific Research on Priority Areas A of "New Protium
Function" and No. 13555182, No. 13875128, No. 12450276, from the Ministry of Education,
Science, Sports and Culture of Japan.
2 Process and Magnetic Properties of Rare-Earth
Bonded Magnets

Jinfang Liu and Michael Walmer

2. 1 Introduction

Bonded magnets can be manufactured by compression molding, extrusion,


injection molding or calendering processes. For compression molding, the
magnet powders coated with epoxies are compression molded into required
geometries and then cured at an appropriate temperature. Bonded magnets
manufactured by this process can be held to tight tolerances, which eliminates
the need for secondary or finish machining. For injection molding, extrusion
and calendering processes, the magnet powders are first blended with
polymers and polymer additives. The mixture is then intensively mixed
together at a certain temperature by a compounder to form a "compound".
The compound can then be extruded, injection-molded or calendered. The
binder that holds the magnet particles together may produce either a flexible or
a rigid bonded magnet. Typical binders for flexible bonded magnets are
elastomers, such as nitrile rubber or vinyl. Typical binders for rigid bonded
magnets include thermoplastics, thermosetting polymers, and in some cases,
metallic binders.
The applications for polymer bonded magnets become increasingly
important in our daily lives. One can find high performance bonded magnets in
various electronic devices, office automation equipment, automotive
components and home appliances, such as computer hard disk drives (HDD),
CDs, DVDs, sensors, etc. Figure 2.1 shows the market share of various
bonded magnets in 1999 (Ring, 2000). According to Ring, the world bonded
magnet market is about $873 mill ion in the year 1999, of which about 42 % are
million
bonded rare earth magnets, 52 % are bonded ferrite (Ring,
(R ing, 2000). Although
bonded ferrite still dominates the market, bonded NdFeB magnets are
becoming more and more important due to their superior magnetic properties.
Figure 2. 2 shows the bonded NdFeB magnet market by appl ication in 1999
application
(R ing, 2000). It can be seen that most of the bonded NdFeB magnets are used
in motor-related applications. Spindle motors of HDDs, CDs and DVDs
account for 55 % of the bonded NdFeB market. Of course there are many other
types of applications that are not represented in Fig. 2.2 due to insignificant
Process and Magnetic Properties of Rare-Earth Bonded Magnets 33

market share as of 1999. The DVD spindle motor application is expected to


increase more than the CD application due to the trend of partial replacement
of CDs by the DVDs.
SmCo

F"'ZJ W?8%
F1"ZJ ~8%
4% NdFeB 1M

NdFeB CM
30%

Figure 2. 1 Market share of various polymer bonded magnets in 1999 (Ring, 2000).

Appliance
Automotive 2%

120/(1~ I\Pi"d1~HDD
120/(\ CjPindle_HDD

""~~
,,~~ ~ 33%
33%
Stepper-OA
31%

. JD
. jt2dleco
Spmdle-DVD
3%
Spmdle-CD
19%

Figure 2.2 Bonded NdFeB magnets by application in 1999 (Ring, 2000).

Industry experts expect an annual revenue growth rate of 16 %, and a


volume growth rate exceeding 25 % .
Following are some advantages and disadvantages comparing polymer
bonded magnets to their sintered counterparts.
Advantages:
( 1) Cost effectiveness: Bonded magnets are made by net-shape or near
(1)
net-shape manufacturing processes, which can be held to tight tolerances
without secondary or finish machining which significantly reduces the
production cost.
(2) Bonded magnets with complex geometries or magnetic assemblies
can be produced using multi-component injection molding techniques.
(3) A broad selection of polymer binders and polymer additives gives the
flexibility for production and meets the requirements of various applications.
34 Jinfang Liu and Michael Walmer

(4) Isotropic bonded magnets can be easily magnetized into various and
complex magnetization patterns.
(5) Fully automated production processes lead to more uniform magnetic
properties and repeatability of precision.
Disadvantages:
(1) Lower magnetic properties compared to their sintered counterparts
due to polymer dilution effect.
(2) The maximum operating temperature is limited, to some extent, by
the temperature characteristics or limitations of polymers.
(3) Higher tooling cost for injection and extrusion molding processes.
For convenience, weight percentage is used in the description of bonded
magnet compositions in this chapter. Because the densities of the polymers
(typically about 1. 0 to 1. 3 gl cm 3 ) and the magnet powders (typically
7.6 gl cm 3 for NdFeB powder and 8.5 gl cm 3 for SmCo powder) are different,
the volume percentage is significantly different from the weight percentage.
The remanence is approximately proportional to the volume percentage (v) of
the magnet powders in the bonded magnets, but not proportional to the weight
percentage. The energy product (BH) max is proportional to v 2 . The volume
percentage should be maximized while maintaining adequate flow
characteristics
characteristics.. Optimization of particle size distribution can help improve the
packing factor, and, therefore the solid loading level. Process and
composition optimization can decrease the void ratio and therefore increase
the density. The density, p, of the bonded magnets can be calculated as
follows: Pp== 1/( wII1 PI + w21 P2 + w31 P3 + w41
11 (WI P4 + ... + wnl
W 4 1P4 W nl Pn)
pn ) where WI'
W 3' W 4' ... , and W n are the weight percentages of magnet powders,
W 2' W3'
W2'
polymers, antioxidants, lubricants, coupling agents and other additives,
=
respectively, and WI + W2 + W3 + W4 + ... + W n = 100% ; PI' P2' P3'
P4' ... , and Pn are the values of density of the magnet powders, polymers,
antioxidants, lubricants, coupling agents and other additives, respectively.

2.2 Recent Progress 3n Rare Earth-Magnetic Powders


for Bonded Magnets

2.2.1
2. 2. 1 Isotropic NdFeB Powders
Isotropic NdFeB powders are the most popular choice for bonded rare earth
magnets. Magnequench International is the largest commercial source of
isotropic NdFeB powders. In the 1980s, Croat and Herbst of General Motors
(Croat et al., 1984; Croat, 1988, 1990, 1994, 1997; Herbst, 1991)
developed a melt-spinning (jet casting) process to produce NdFeB-based
Process and Magnetic Properties of Rare-Earth Bonded Magnets 35

isotrooic powders. In the melt-spinning process, a stream of molten alloy is


directed onto the outer surface of a rapidly spinning wheel to form thin ribbons
or flakes. The average quench rate is estimated to be around 105 K/s. An
inert gas is used to protect the molten alloys and ribbons from oxidation at
elevated temperatures. Because of the high cool cooling
ing rate, free iron
precipitation is avoided, and therefore, the composition of melt-spun alloys is
almost stoichiometric. The magnetic properties of the melt-spun ribbons
depend upon the alloy composition: quench rate, molten alloy temperature and
melt-spinning conditions, etc. Because of the narrow window of quench rate
required for optimized magnetic properties, the common practice in the
industry is to overquench the molten alloys at a high cooling rate to produce
amorphous ribbons, which are annealed afterwards to achieve the optimum
magnetic properties. Isotropic NdFeB powders/flakes produced by the melt-
spinning/jet-casting process have attracted considerable attention both from
the magnetics industry and academia (Manaf et al., 1991; Hadjipanayis and
Krause 1995; Dahlgren et ai., aI., 1997a, 1997b; Goll et ai.,aI., 1998). Research
work has been focused on the improvement of magnetic properties
(Hadjipanayis and Gong, 1988; Liu et ai., al., 1994a), microstructure-property
relationships (Davies et ai., al., 1996) and coercivity mechanism (Liu et al. ,
1994b, 1994c) of nanocrystalline
nanocrystall ine magnetic materials.
Table 2. 1 shows the typical magnetic properties of Magnequench
isotropic NdFeB powders (Magnequench, 2000a). Magnetic properties of
isotropic NdFeB powders are normally measured using a vibrating sample
magnetometer (VSM). VSM measurement gives the open circuit results, thus
the self-demagnetization of the NdFeB powders needs to be considered. The
NdFeB powders have irregular shapes; therefore it is almost impossible to
accurately calculate their demagnetization factor (N). Herchenroeder
( Herchenroeder, 2000) used VSM to experimentally determine the
demagnetization factor of the isotropic NdFeB powder. The magnetic field H = =
H.a - H d = H.
H H a - NM, where H H.a is the applied field, H d is the self-
demagnetization field, N (0< N < 1) is the demagnetization factor. The first
quadrant magnetization for high permeability magnetic materials at fields near
zero is MM== aHa = exa ( HH + HHd) == aex ( MM// k + NM), where aex is the apparent
exH. =
magnetic susceptibility, k is the true magnetic susceptibility. Therefore, N= N=
l/a-l/k.
1/ex-1/k.
Henchenroeder (Herchenroeder, 2000) calculated a ex and k values on
overquenched NdFeB powders/flakes using the VSM measurement data, and
the demagnetization factor, N, is then determ ined as 0.21
determined O. 21 for Magnequench
NdFeB isotropic powders. The magnetic properties in Table 2. 1 are VSM test
= O. 21. It should be pointed
results corrected by a demagnetization factor N =
out that the crushed melt-spun NdFeB powders/flakes have different
demagnetization factors if the manufacturing process is different. Melt-spun
36 Jinfang Liu and Michael Walmer

ribbons were measured in the plane and perpendicular to the spinning direction
(i. e. , across the width) by some researchers (Liu and Davies, 1996; Davies
et al., 1996). No correction for self-demagnetization is needed in this case
since the ribbon thickness, t r' is typically a factor of 30 to 100 times smaller
than the ribbon width. The most popular isotropic NdFeB powder so far is
MOP-B
MQP-B due to its ease of magnetizing and good magnetic properties. It is a
popular choice for multi-pole stepper and spindle motor applications. MOP-A
MQP-A
powder has a higher coercivity and higher resistance to demagnetization;
MOP-C MOP-A but with a more stable reversible temperature
MQP-C is similar to MQP-A
coefficient; MQP-D
MOP-D has a similar temperature coefficient with MOP-C,
MQP-C, but with
lower coercivity and a moderate magnetization requirement; MOP-O MQP-O is
designed for higher temperature applications; MOP-B MQP-B + has a maximum
energy product of 130 kJ/m 3 ((16. 16. 3 MGOe), the highest among all
Magnequench, powders (Magnequench, 2000a).

Table 2. 1 Magnetic properties of isotropic NdFeB powders.


Maximum Energy
Residual Induction Intrinsic Coercivity Coercive Force
Isotropic Product
B,r
B jH c
iHc He
Powder BH)ma<
( BH)m.,
(kG)
(kG> (mT) (MGOe) (kJ/m 3 ) (kOe) (kA!m) (kOe) (kA/m)

MOP-B 88 880 14.9


149 119 9.0
90 720 64 510
MOP-D 87 870 153 121 10.3
103 815 6.6
66 525
MOP-B+ 9 1 910 16.3 130 98 775 68 540
MOP-O 80 805 133 106 12.5
125 995 66 525
MOP-A 80 800 13 1 104 150 1190 6.5
65 515
MOP-C 8.0 800 13.2
132 105 16.5
165 1310 6.6
66 520

Typical magnetic properties of bonded NdFeB magnets produced by


Electron Energy Corporation are shown in Table 2. 2 (Electron, 2000).
Typical maximum energy product (BH) max is 80, 64 and 48 kJ/m 3 for
compression, extrusion and injection molded isotropic bonded NdFeB
magnets, respectively. Figure 2. 3 shows the typical demagnetization curves
of polymer-bonded NdFeB magnets, where curves 1, 2, 3, 4, and 5 are the
second quadrant demagnetization curves of EEC NEOHS9, NEOH9, NEOT9,
NEOlS 10, and NEOll 0, respectively (Electron, 2000). Figure 2.4 shows the
typical demagnetization curves of compression-molded MOl
MQl bonded NdFeB
magnets between - 40 to 150 'c (Magnequench, 2000a).
Process and Magnetic Properties of Rare-Earth Bonded Magnets 37

Table 2. 2 Typical magnetic properties of isotropic bonded NdFeB magnets.

Bonded Residual induction, Coercive force, Intrinsic coercive Maximum energy


T~)
Ti,J,'
magnet B, He force, ,He
;H e product, (BH) max
grade
(kG) (mT) (kOe) (kA!m) (kOe) (kA/m)
(kA!m) (MGOe) (kJ/m3 ) ('C)
eC)

NEOl6
NEOL6 5.0-5.6 500 - 560 4.1-4.5 326
326-- 358 8-10 637 -796 5.5-6.5 44-52 120

NEOL8
NEOl8 6.0-6.5 600
600- 650 4.7-5.0 374 - 398
-650 8-10 637 -796 7.5-8.5 60-68 120

NEOLlO
NE0110 6.8-7.4 680-740
680 -740 5.1-5.4 406-
406 - 430 8-10 637 -796 9.5-10.5 76-84 120

NEOLS6
NEOlS6 5.0-5.5 500
500-- 550 4.2-45
4.2-4.5 334 - 358 9-12 716 - 955 55-65
5 5-65 44-52
44 -52 130

NEOLS8
NEOlS8 5.9-6.4
5.9 - 6 4 590
590- 4.8-5.2 382 - 414
- 640 4.8-52 9-12 716-955 7.5-8.5 60-68 130

NEOlSIO
NEOLS10 67-7.2
6.7-7.2 670 -720 5.4-5.8 430-
430 - 462 9-12 716-955
716 -955 95-105 76-84 130

NEOH5 4.3-4.8 430-


430 - 480 4.0-4.4
4 0- 4 4 318-
318 - 350 13-17 1034 --1353
1353 45-5
4 5- 5 5 36-44 130

NEOH7 52-5.6 520 - 560 44.7-5.1


5.2-5.6 520-560 7-5 1 374 - 406 13-17 1034 -1353 6.5-75
65-7.5 52-60 130

NEOH9 6.0-6.5 600-


600 -650
650 54-5.9
5.4-5.9 430 - 470 13-17 1034 - 1353 85-9.5
8.5-9.5 68-76 130

NEOHS5 42-4.8
4.2-4.8 420 -480
- 480 3.9-4.5
3 9-4 5 310- 358 14-18 1114-1433 4.5-5.5 36-44 130

NEOHS7 5.2-5.7 520-


520 - 570 4.7-5.2 374-414 14-18 1114-1433 6.5-7.5 52-60 130

NEOHS9 6.0-6.4 600-


600-640
640 5.3-5.8 422-
422 - 462 14-18 1114-1433 8.5-9.5 68-76 130

NEOT5 4.4-4.9 440- 490 3.9-4.4 310-350 11-14 876-1114


876 -1114 4.5-5.5 36-44 150

NEOT7 5.3-5.8 530-


530 - 580 4.5-5.2 358-
358 - 414 11-14 876-1114 6.5-7.5 52-60 150

NEOT9 6.1-66 610 - 660 5.0-5.8


6.1-6.6 610-660 5.0-5.6 398 - 446 11-14 876-1114
876 -1114 8.5-9.5 68-76 150
(1)
( 1) Top is the maximum operating temperature, which is dependent upon permeance coefficient,
coating and environment.

800 ,-----.-----,----,.-------,
,------,----,-----,--------,

600 f---+-----+---,;>o~.'!!fs:~1'IJI
600 I-I--I----=:;~~~~
f='
i='
5~
'<:i
CQ 400 i-----t---/-F+-+-:---f--t----i'fH----i
f------+---/--F-+--r.--f--t-----l'fH----i
....
....,o

f----A--I-:--I--H-I'------f-I/I----I
200 1------t1--f:-+-tH'------f-fll----I

O'--_.L-----'-...L---L-_...J..L_
O'--_-'------'-...L---L-_--'-'-_ _-""'--_ _---'
_--'
-1600 -1200 -800 -400
H(kA/m)

Figure 2. 3 Typical demagnetization curves of NdFeB isotropic bonded magnets.


Curves 1, 2, 3, 4, and 5 are the second quadrant demagnetization curves of EEC
NEOHS9, NEOH9, NEOT9, NEOLS10,
NEOlS 10, and NEOL
NEOl110,
0, respectively (Electron, 2000).
38 Jinfang Liu and Michael Walmer

8,-------.---,---------r--.,---,-------,
8 ,----------r--~-___.--_,_--,_____-__, 800
7 700
~ 6 f------+--+----t--=-'"'f'C---=~j__~"7'I
f----+--+----+--=--F----=~\--~"7'I 600

g5 PE
500 f='
Cl:l 4
'l:l 400 ;::
o(; ...
......
~ 3 300 0
(;
~
~
~22 200
~
~ ...,

I\ f------++--+-/----t-f----k-I'----F-I-j__---l 100
\00
oo'----_-"-__
'---_-lL_ _ L..L-_--'--'-..LJ.L.L...L---'-_--.J
L..J-_-L.-L..-L...LL.L--'--L._----'

-\200
-1200 -1000 -800 -600 -400 -200 0
H(kA/m)

Figure 2.4 Demagnetization curves of MOl bonded magnet made from MOP-B powder
(Magnequench, 2000).

2. 2. 2 Lean Neo Isotropic Powders


In the 1980s, Coehoorn and coworkers reported an enhanced remanence
(M,/ M s ) in nanocrystalline Fe3 B/Nd2 2 Fe14 B composite (Coehoorn et al. ,

1988, 1989). Magnequench developed the MOP-O MOP-Q composition, which is


based upon ex-Fe/ Nd2FeI4B.
Nd2 Fe14 B. There is an exchange coupling between the soft
and hard magnetic phases in these materials. These magnets are often
referred to as exchange-coupled magnets, exchange-spring magnets, or
nanocomposite magnets. In the magnetics industry, Fe3 B/Nd2 2 Fe14 B or ex-Fe/

Nd22 Fe14 B magnets are also called "lean neo" magnets. The magnetic
hardening mechanism of nanocomposite magnets has been discussed by
Kneller and Howig (Kneller and Howig, 1991). Detailed information on
nanocomposite magnets can be found in other chapters of this book. The
desired microstructure of nanocomposite magnets consists of nano-sized grains
of hard and soft magnetic phases, typically 10 to 20 nm. Nanocomposite
powders are normally produced by melt spinning, followed by annealing,anneal ing, to
achieve nanocrystalline
nanocrystall ine microstructure. Although remanence B B,r values as high
as 1. 2 T have been reported for this type of materials (Croat, 1997),
nanocomposite magnetic powder with comparable B, is not commercially
available as of this writing. Lean neo powder MOP-Q,MOP-O, which is commercially
available from Magnequench International, has a B rr of 0.86 O. 86 T. The typical
MQP-Q powder are shown in Table 2. 3. Figure 2. 5
magnetic properties of MOP-O
shows the typical demagnetization curves of lean neo bonded magnets using
the commercial available powder. Due to their low intrinsic coercivity, the
lean neo materials cannot be used at elevated temperatures. Low coercivity
lean neo materials can be easily magnetized, which is an advantage in
designing smaller stepper motors for modern consumer electronic and
appl ications. According to Croat, the number of functions
computer peripheral applications.
now available on modern consumer electronic products has resulted in the
Process and Magnetic Properties of Rare-Earth Bonded Magnets 39

production of a large number of extremely small stepper motors with rotors as


small as 4 - 6 mm with eight-pole magnetization. The lean neo material can be
magnetized up to 80 % while the single-phase material can only be magnetized
to about 55 % of saturation under a maximum magnetization field of 800 to
860 kA/m (Croat, 1997). The good magnetizability of the lean neo magnet
may lead to more new applications in the future because of the trend toward
miniaturization in the electronics industry.

Table 2.3
2. 3 Typical magnetic properties of MQP-Q powder and its bonded magnet.
Residual induction, Coercive force, Intrinsic coercive Maximum energy
Bonded
magnet
B, He force,; He product, (BH) max
grade (kG) (mT) (kOe) (kA/m) (kOe) (kA/m) (MGOe) (kJ/m 3 )

MQP-Q
8.2-8.8 820
820- 2.5-3.5 200
- 880 25-3.5 200- 3.5-50 280- 400 6 8-10
- 280 3.5-5.0 54-80
powder

MQP-Q
bonded 65-6.9 650-690
650- 690 >2.5 >200 >35 >280 >5.5
>55 >44
magnet

H(kOe)
-7.5 -5.0 -2.5 o
8 800
7 700
~ 600
~6
g5
o ~V
./- / ./
65 V
./ 500 i='
Cl:l 44 ./'" './
.-/ / .s
400 ...
::
oo -40C / /
-40'C / 300 ...,
'-,
l~ ~ /
/ / /
/20'C /IOO'C
/20C /lOOC
200
I 100
o // I/
-600 -400 -200 o
H(kA/m)

Figure 2. 5 Typical demagnetization curves of lean neo bonded magnets


(Magnequench, 2000).

2. 2. 3 Anisotropic NdFeB Powders for Bonded Magnets


Anisotropic NdFeB powders can be produced by hydrogenation-
disproportionation-desorption-recombination (HDDR) process (Takeshita and
Nakayama, 1989; Takeshita and Morimoto, 1996; Nakayama et ai., aI., 1991;
Nakayama and Takeshita 1993; Uehara et ai., al. , 1993; Matzinger et al. , 1995;
aI., 1995). Detailed
Ahmed et al., Detai led information on this process can be found in
other chapters of this book. Maximum energy products range from 15 to 18
MGOe for compression molded anisotropic epoxy bonded NdFeB magnets. The
40 Jinfang Liu and Michael Walmer

temperature coefficient of intrinsic coercivity of these bonded magnets is


O. 55 % FC , which is larger than that of the isotropic bonded NdFeB
about - 0.55
O. 45 % FC) but smaller than that of the sintered NdFeB
magnets (typically - 0.45
magnets (typically - 0.65 % FC) (Croat, 1997). The typical demagnetization
curve of the bonded anisotropic NdFeB magnets is shown in Fig. 2. 6 (Liu,
1999a). The typical magnetic characteristics are listed in Table 2.4 (Croat,
1997; Liu, 1999a).

Table 2.4
2. 4 Properties of compression molded anisotropic bonded NdFeB magnets.
Magnetic properties CG8 81

Residual induction, B, 8. 1-8 5 kG


8 810-850 mT

Coercive force, He 6.
6 5-7
5 -7 5 kOe 520 - 600 kA/m

Intrinsic coercivity, ;i He 11. 5 - 14 . 5 kOe


115-14 920-
920 - 1160 kA/m

Maximum energy product, (BH)m.,


(BH)ma, 14- 17 MGOe
14-17 112- 136 kJ/m 3
112-136

Required magnetizing force, H


H,s 30 kOe 2400 kA/m

Temperature Coefficient of B, to 100 c


'c - o. 09%/'C
Temperature coefficient of i He to 100 c
'c -0.55%I'C
-0.55%/,C

1000
~
....-J
800

P
f=" 600
V
/ /
5S
~ 400
';:
/ /
o
~
'-,

200 1/II /
o
-1280 -960 --640
640
/ -320
H(kA/m)

Figure 2. 6 Demagnetization curve of compression molded anisotropic NdFeB


magnets (Liu, 1999a).

2. 2. 4 SmCo Powder for Bonded Magnet Applications


Sm(Co,
Sm( Co, Cu, Fe, Zr) Zr)zz permanent magnets were developed in the 1960s and
1970s (Nesbitt et al. , 1968; Ray and Strnat, 1972; Perry and Menth, 1975;
OJ ima et al. , 1977; Ray, 1984).
Ojima
SmCo sintered magnets have attracted more attention due to high
operating temperature requirements for defense applications (Liu et al ..
Process and Magnetic Properties of Rare-Earth Bonded Magnets 41

1998a, 1998b, 1999b, 1999c, 1999d 199ge; Liu and Hadj ipanayis, 1999;
Zhang et al. , 2000; Chen et al. , 1998,2000; 1987>. SmCo anisotropic
1998, 2000; Kim, 1987).
powders for bonded magnets have also been developed to meet the higher
operating-temperature requirements for demanding appl ications. The
production routes for bonded SmCo magnets are summarized in Fig. 2.7. 2. 7. The
cast Sm (CCo,
Co, Cu, Fe, Zr) z ingots are normally solution treated at 1120 to
1180 'c for 5 to 20 h, and then aged at 820 to 880 'c for 3 to 10 h followed by
a slow cooling to 400 'c at a rate of 1 to 2 'c per minute. The fully processed
ingot is then pulverized to a powder with particle size less than 200 IJm. The
powder is finally tested before compounding or coating. Okonogi and Shimoda
(Okonogi
COkonogi and Shimoda, 1985) claimed that formation of columnar grains
during casting is critical for producing high-quality SmCo powder for bonded
magnets. Figure 2.8 is a sketch of the cross-sectional view of the SmCo ingot
(Okonogi
COkonogi and Shimoda 1985).

Mixing elements according to the


predetermined formula

Induction melting and cast into a mold to get


favorable columnar structure

Solution treatment followed by quenching

Heat treatment at 800 to 870 'C


followed by a slow cooling

Crushing and particle size adjustment

Coating or blending powder with epoxies


and polymer additives

Magnetic and mechanical inspection

Packing and shipping

Figure 2.7
2. 7 Production routes for bonded SmCo magnets.
42 Jinfang Liu and Michael Walmer

Figure 2.8 A sketch of the cross-sectional view of the SmCo ingot, where M is the
l<igure
mold, A is the surface area of the ingot, C is mostly columnar grains, and E is
equiaxed grains in the center of the ingot (Okonogi and Shimoda, 1985).

As we know, the solidification of molten alloys starts from the cold wall of
the mold. The crystals formed near the mold-wall will grow into the liquid alloy
while competing with neighboring crystals. Region A in Fig. 2.8 is the surface
area of the ingot. Next to the surface area is the columnar zone, which is
formed due to the preferred growth of the crystals along the direction parallel
to the heat gradient. Zone E is the center of the ingot with equiaxed grains.
According to Okonogi and Shimoda (1985), the percentage of columnar zone
within the ingot can be increased by properly controlling the temperature of the
molten alloys before casting, the manner of casting, and the cooling rate of the
molten alloys. They further stated that the melting temperature should be at
least 320 C above the melting point of the alloy composition in order to
achieve desired columnar structure. Figures 2. 9 and 2. 10 are the typical
demagnetization curves of high coercivity and low coercivity bonded SmCo
magnets, respectively (Electron, 2000). The coercivity of the bonded SmCo
magnets depends upon the alloy composition as well as the casting, solution
treatment and ageing conditions of the alloy. Table 2. 5 lists the magnetic
properties of bonded SmCo magnets produced at Electron Energy Corporation
(Electron, 2000). Epoxies, nylon 12 and polyphenylene sulfide (PPS) are the
preferred binders for bonded SmCo magnets. The PPS bonded high coercivity
SmCo magnets can be used at temperatures up to 250 C .

Table 2.5
2. 5 Magnetic properties of bonded SmCo magnets.

Bonded Residual induction, Coercive force, Intrinsic coercive Maximum energy


Top
magnet B, He force, ;,He
He product, (BH) max
grade (kG)
(kG> (mT) (kOe) (kA!m) (kOe) (kA/m) (MGOe) (kJ/m 3 ) ( 'C)

SmCo H5 4 3- 5 0 430
430-- 500 44.0-4.6
0-4 6 318 - 366 >20 >1592 4.5-55 36-44 250
SmCo H10 66.1-
1- 6. 8 610-680
610 - 680 5 5-6 2 438 - 494 >20 >1592 9-11 72-88
72- 88 250
SmCo L5 44-5.0
4 4- 5 0 440 - 500 8-44 4
3 8- 302 - 350 >12 >955 4 5- 5 5 36- 44 200
L10 6.3-7.0
SmCo Ll0 6 3-7 0 630 -700 5 2- 5 8 414-462 >12 >955 9-11 72- 88 200

SmCo L14 7 8- 8 4 780-


780 - 840 6 1-6.8
6.1-6.8 486-
486 - 541 >12 >955 13-15 103- 119 200
Process and Magnetic Properties of Rare-Earth Bonded Magnets 43

H(kA/m)
-1600 -1200 -800 -400
8

2fL----/--+---f----Ic'----+1J'III--\---'H1200
2fL---/--+---f---I:'--------Hi'IH--\--'H-l200

o '---L _ _- -
'------L -'--'---_
_----'-
-'---
-- - _ _--.lL""-_
-' -_ _-'"
--ILWL-_ _- '"

-20 -15 -10 -5 0


H(kOe)

Figure 2.9 Typical demagnetization curves of high coercivity bonded SmCo magnets at
various temperatures (Electron, 2000).

H(kNm)
H(kA/m)
-1200 -800 -400 0
10 1000

8 800
G
6' E='
f='
c6 6 600
OQ
Cl:l
....
55-
OQ
Cl:l
0 ....
~ 4 400 ....,
0
~
'7
"""
2 200

0
-15 0
H(kOe)

Figure 2. 10 Typical demagnetization curves of low coercivity bonded SmCo magnets at


various temperatures (Electron, 2000).

2.2.5
2. 2. 5 Other Magnetic Powders
Some new magnet alloys currently under development are possible candidates
for bonded magnet applications. One of the promising candidates is SmFeN
magnet powders. SmFe alloys can be produced by either reduction-diffusion
process or induction melting. The nitrogenation of SmFe alloy powders leads
to the formation of Sm2 Fell Nx anisotropic coercive powder.
powder, which is then
pulverized to about 2 IJm followed by surface treatment and mixing with
binders. The powder mixtures or compounds can then be used for bonded
44 Jinfang Liu and Michael Walmer

magnet production.
Ferrite is so far still the most widely used powder for bonded magnets.
Bonded ferrite is the least expensive option with quite good thermal stability
and excellent corrosion resistance. But the maximum energy product is
normally below 16 kJ/m 3 . Hard ferrite powders are based on the phases of
BaFel2 0 19 19 or SrFe12
SrFel2 0 19
19 as well as their solid solutions, for example
Bal- x Sr x Fel2 0 19 , Barium or Strontium carbonates, iron oxides and minor
Bal-xSrxFe12019'
additives are first weighed and then blended. The blended powders are
processed using a calcining furnace where they are chemically reacted into
barium or strontium hexaferrites. The calcined powders are treated to achieve
a suitable particle size and shape. The ferrite powder is then compounded and
is then ready for bonded magnet production. Table 2. 6 lists the typical
magnetic properties of bonded ferrite (Ormerod and Constantinides, 1997).
Ferrite can also be blended with other powders, such as lean neo powder, to
make a "hybrid" bonded magnet. Because the temperature coefficient of
intrinsic coercivity is positive for ferrite but negative for lean neo, the hybrid
magnet with about 20 weight percent lean neo can have an almost zero
temperature coefficient of intrinsic coercivity (Ormerod and Constantinides,
1997)
1997).. For more information on ferrite materials, please refer to
Ferromagnetic Materials edited by E. P. Wohlfarth (Wohlfarth, 1982) and
the references therein.

Table 2.6 Typical magnetic properties of bonded ferrite.


Isotropic Anisotropic
Manufacturing method
CGS(MGOe)
CGSCMGOe) SICkJ/m 3 )
SI(kJ/m CGS(MGOe)
CGSCMGOe) SICkJ/m 3 )
SI(kJ/m

Calendered ferrite o 5-0 7


0.5-0.7 4-5 6 1.4-1 6 11.1-12.7

Extruded ferrite 0.4-06


04-0.6 3.2-4.8
3 2-4 8 1.2-1 5 96-11.9
9.6- 11.9

Injection molded ferrite o 5- 0 8


05-0.8 4.0-6.4
4 0-6 4 1 5-1.77
1.5-1 119-13.5
11.9-135

2. 3 Process and Magnetic Properties of Polymer Bonded


Rare Earth Magnets

2.3.1 Bonded Magnets Produced by Compression Molding Process


Bonded rare earth magnets and bonded Ferrite are the two basic groups of
bonded magnets, although there are small amounts of bonded Alnico magnets
and other hybrids on the market.
In 1999, the bonded rare earth magnets accounted for about 42 % of the
Process and Magnetic Properties of Rare-Earth Bonded Magnets 45

entire bonded magnet market, of which 71. 5 % are compression molded


NdFeB magnets, 19 % are injection molded NdFeB magnets, 9. 5 % are
bonded SmCo magnets produced by all means (Ring, 2000).
It should be noted that the majority of bonded rare earth magnets
(71. 5 %) are compression molded, therefore our emphasis in this chapter will
be on compression molding. Injection molding and extrusion process for
bonded rare earth magnets will also be briefly discussed.
Figure 2. 11 is a sketch of a compression molding machine, which is also
called a powder compacting press. In Fig. 2. 11 , A and B are upper and lower
punches, respectively; C is the magnetizing coil; 0 is the epoxy-coated
magnetic powders or other powder mixtures ready to be compacted to bonded
magnets with cylindrical shape. Tooling design varies with the required
geometry. Both mechanical and hydraulic presses are available on the
market. Mechanical presses normally have low horsepower requirements, a
wide range of pressing forces and high production rates, while hydraulic
presses have greater depth of fill and overload protection. The binder for
compression-molded magnets is normally thermosetting epoxies. The majority
of bonded neodymium-iron-boron magnets to date are produced by
compression molding. The advantage of compression molding is high solid
loading, and therefore high maximum energy product (BH) max' In the case of
98.5 %
the epoxy bonded NdFeB magnets, the solid loading can be as high as 98.5%
by weight. But compression molding is only suitable for simpler geometries,
such as discs, rings, rectangles, and arcs. There are many useful features in
the compacting presses. The presses can be designed in either a part-height

Figure 2. 11 Sketch showing a compression-molding machine, where A and B


are the upper and lower punches, respectively; C is the magnetizing coil, D
0 is
the magnetic powder mixtures ready for compression molding.
46 Jinfang Liu and Michael Walmer

dependent mode or in a pressure dependent mode. In a part-height dependent


mode, production parts with consistent part-height can be produced. When the
part-height is reached, compaction is stopped immediately. Or, if pressing
force is reached and the pressure dwell timer times out, compaction is
stopped. The press will normally go to whatever pressing force is required to
get the target part height unless the maximum pressure is reached first. In a
pressure dependent mode, parts with consistent density can be produced.
Most of the well-equipped automatic presses will automatically adjust the fill
stroke in order to maintain the target compact height or pressing force. When
the press is in the pressure dependent mode, the fi II is increased or decreased
in order to keep the part height on its target. When the press is in the part-
height dependent mode, the fill position is modified in order to keep the
pressing force on its target. Modern presses normally have computerized
control panels, and all control parameters are digital. There are many other
useful features, for example, pre-press pause, die-force float, and fill
compensation. One can find more information on compression molding
technology in the book entitled Powder Metallurgy Equipment Manual by
Samuel Bradbury (Bradbury, 1986).
Magnet particles are generally coated with thermosetting epoxies. The
qual ity of epoxies is judged by their adhesion, impact strength, hardness,
quality
thermal shock, abrasion resistance, penetration and chemical resistance. The
coating of magnetic particles has the following advantages:
( 1) The magnetic particles will be ready for pressing after coating;
(1)
(2) Coated particles have improved powder flow characteristics;
(3) Coated particles will improve the uniformity of the molded parts;
(4) Higher production rate are possible;
(5) Higher density can be achieved with a relatively low pressure;
(6) Higher solid loading level is possible;
(7) Coated particles have improved corrosion protection.
The coated particles can be compression molded into certain geometries.
The molded parts are cured in an oven at a pre-determined temperature. For
high volume production, the molded parts are transferred automatically into a
curing oven with conveyors; the cured parts are then cleansed before going to
the epoxy coating line to coat the surface of the magnets. The surface coating
of bonded magnets can be done either by electrostatic coating or spray coating
process. Electrostatic coating, also known as e-coat, is accomplished by
charging particles with high voltage as they are sprayed onto bonded magnets
at ground potential. Electrostatic coating has the following advantages:
(1) Excellent finish quality;
qual ity;
(2) High productivity;
(3) Cost effectiveness;
(4) Can coat parts with complex geometries;
(5) Tight tolerance.
This coating method is ideal for flat, angular, or irregular shaped objects. The
Process and Magnetic Properties of Rare-Earth Bonded Magnets 47

process characteristics of e-coat and spray coating are listed in Tables 2. 7


and 2. 8, respectively. The surface coated magnets are inspected both
magnetically and mechanically before packaging and shipping. Epoxy surface
coating is recommended for all compression molded NdFeB magnets in order to
protect the magnets from corrosion. Surface epoxy coating is generally not
necessary for bonded SmCo magnets. SmCo material has excellent corrosion
resistance.
res istance.

Table 2.7 Process characteristics of e-coat.


Application method Immersion electrodeposition
epoxy/urethane water based

Pre-treat process options Alkaline clean


Acid etch/passivate

Thickness 15-25 ~m

Pencil hardness 2H-4H

Table 2.8 Spray coating characteristics.


Material Epoxy, phenolic,
phenol ic, and acrylic
acryl ic

Thickness 20 ~m nominal

Acid resistance Good

Pencil hardness up to 5H
Up

Operating temperature -73-204


-73 - 204 'C

Compression molded magnets have certain limitations on the geometry,


dimensional, and mechanical tolerances. The common shapes of compression
molded NdFeB magnets are blocks, discs, rings, and arcs. Table 2.9
contains the design and manufacturing guidelines for compression-molded
NdFeB magnets as recommended by Magnequench International
CMagnequench, 2000b). The tolerances Iisted
(Magnequench, listed in Table 2.9
2. 9 apply to uncoated
or impregnated bonded magnets without grinding. Due to the thickness of the
coating, O.
0.012
012 mm per surface should be added to the tolerances for coated
magnets. Better tolerances can be achieved by the careful control of the
production process and tight-tolerance tooling design.
The maximum energy product, CBH)max'
(BH) max' of compression molded isotropic
NdFeB magnets is about 10 MGOe. Some producers claim they can produce
bonded isotropic NdFeB magnets up to 12 MGOe. The intrinsic coercivity of
compression-molded magnets varies from 4 to 17 kOe depending upon the
material grades. The temperature coefficient of residual induction, B B"r' ranges
from - O. 11 % FC to - O. 07 % FC. The temperature coefficient of intrinsic
coercivity i He for bonded NdFeB magnets is about - 0.4 % FC , which is much
smaller than that for their sintered counterparts ( C-- O. 65 % FC ). High
48 Jinfang Liu and Michael Walmer

temperature performance is always a concern for bonded NdFeB magnets.

Table 2.9 Design and manufacturing guidelines for compression molded NdFeB magnets.

Shape of compression Dimensions or


Description
molded magnets tolerance

T ,L, or W
Smallest dimension, T,L,orW 1 .52 mm
1.52

... Maximum thickness, T 25.4 mm

~
Rectangular
h magnets Tightest tolerance on thickness, T O. 13 mm
pressed along
'y
\J dimension T Tightest tolerance on the width, W, and
0.1O mm
length, L

Minimum die corner radii 0.51 mm

Maximum diameter, 0 60.8 mm

Minimum diameter, 0 3.30 mm

W
Discs or round

bd j
1 bars pressed
Tightest tolerance on the diameter, 0 O. 10 mm
along the
dimension H Maximum height, H 25.4 mm

Minimum height, H 1.52 mm

Tightest to lerance on part height, H O 13 mm

Maximum 00 152.
152 4 mm

Minimum 00 6.35
6 35 mm
OD Minimum 10 2. o mm
2.0
I .
I ID .. II

oc
(l:: ~.
Minimum wall thickness 1 mm

::t::
:I:: Maximum height, H 25.4 mm

....... -/"-
- Minimum height, H 1.52 mm
"-- .

Tightest tolerance on the part height, H O. 15 mm

Maximum cross section area 29 cm 2

Previously, bonded NdFeB magnets can only be used up to 110 'C.


Through chemical composition modification, the high temperature
characteristics have been significantly improved in recent years. Substitution
of cobalt for iron improves the high temperature capability, but the cobalt
market pricing is volatile. Nb and Zr additions are also found to be effective in
improving high temperature capability. Some of the isotropic bonded NdFeB
magnets can now be used at temperatures up to 150 to 180 'c in certain
magnetic circuits.
Process and Magnetic Properties of Rare-Earth Bonded Magnets 49

2. 3. 2 Hot Compression Molding Process


Hot compression molding offers some advantages over the conventional
compression molding process. Higher density can be achieved by hot
compression molding due to the improved flow ability of the polymers. In the
case of anisotropic bonded magnets, better alignment can be achieved at
elevated temperatures with the same magnetizing field. For bonded magnets
using some of the thermosetting resins as binders, a separate curing process
may not be needed. A hot press was built at Electron Energy Corporation to do
experiments on compression-molded magnets using thermoplastic and
thermosetting resins. The hot press consists of a hydraulic press system,
frames, a transverse magnetizing coil, a die with heaters, upper and lower
punches, thermocouples, and a temperature controller. The following is a
summary of the experimental results:
( 1) An extruder is used to make the compound with the following
composition:
96wt. % MQP-B
O. 2wt. % zinc stearate
O. 6wt. % antioxidant
3.2wt.
3. 2wt. % Nylon 12.
The above compound was granulated and fed into the die cavity. It was then
compression molded under various pressures at 25, 120 and 210C, 210 'c,
respectively. It was found that hot compression molding of the above
210 'c could produce samples with a density of 5.7 g/cm
composition at 210C gfcm 3 using
2
only 1 tf
t/cm
cm pressure. By comparison, the pressing force would need to be at
cm 2 if pressing at room temperature. Figure 2. 12 shows the pressure
tfcm
least 8 t/
dependence of density of bonded magnets pressed at various temperatures.
120 'c does not have a significant effect on density, because
Pressing at 120C

5.6 f---D--==------I--~---+-___l
f-----o--""""''r'-~__\~~-+~~-+-__\

5.4 1----+---+--~)=:-___7!l~--+_I
1---+-~~+-~---7.:"""'-:~-.Al~~-+-----1

~
~

E
~ 5.2

.~ 5.0 f----+-----/--.,.<-j-
f------+--~+-J- -
--0-
0-Pressed at 120C
120'C
cc:
v
(J)
-
- -0Pressed at 210 C
0-- 'C
a I---(h~--j-
Cl 4.8 f---(j--,I-------j- _ _ Pressed at 25 'C

4.6 L--L.._ _--'----_ _L - _ - - - - I ._ _----'--------.J


L----'---~~--'-----~~L-~----..L~~--L___J

2 3 4 5
Pressure (t/cm 2)

Figure 2. 12 Pressing pressure and temperature dependence of the density of


thermoplastic bonded NdFeB magnet.
50 Jinfang Liu and Michael Walmer

120 'c is below the melting point of Nylon 12. When the samples were pressed
at 210 'c, the nylon binder melted and flowed easily even under a low pressure.
(2) Liquid coated MQLP-B powder
Liquid coated MQLP-B powder with trace lubricant was compression
var ious pressures at 25 and 160 'C. The pressure and
molded under various
temperature dependence of density is shown in Fig. 2. 13. By compression
molding at 160 'c, higher density and magnetic properties were obtained at
lower pressures when compared to room temperature compression molding.
The difference in density between pressing at 160 and 25 'c is more apparent
at lower pressures.
6.5

6.0
11 160 c
Pressed at 160'C l---'
l.----"
t,...--
t,...---
~
t/
~
~
~

Eu
29 5.5
~
./
/
oC- V V
.;;;
.u;
"'"c
<l)
c
25'C
hressed at 25
o 5.0
1/
4.5
o 2 4 6 8 10 12
Pressure (t/cm 2 )
Figure 2. 13 Pressing pressure and temperature dependence of the density of
epoxy-bonded NdFeB magnets.

(3) Bonded anisotropic NdFeB magnets


The compression molded anisotropic NdFeB magnets made from HDDR HOOR
powder have a maximum energy product as high as 18 MGOe. The
temperature coefficient of intrinsic coercivity of anisotropic bonded NdFeB
magnets is about - 0.55 % FC
Fc , which is much larger than that of the isotropic
bonded NdFeB magnets. The following compositions were used in the
experiments:
Composition A:
HODR NdFeB powder
97wt. % HDDR
2. 5wt. % nylon 12
O. 3wt. % antioxidants
O. 2wt. % lubricant and trace other additives.
Composition B:
HOOR NdFeB powder
97wt. % HDDR
2. 5wt. % polyphenylene sulfide
O. 3wt. % antioxidants
0.2wt. % lubricant and trace other additives.
The above compositions were compression molded using a transverse hot
Process and Magnetic Properties of Rare-Earth Bonded Magnets 51

press. The al alignment


ignment field is 1114 kA/m
kA/ m C14 kOe). Compression pressures
were varied from 1 to 6 t/ cm 2 2
. The compression molding temperature and

pressure dependence of maximum energy products is summarized in Table


2. 10. Samples POl PO 1 to P04 have relatively higher CBH ))max
max compared to

samples NO 1 to N04 under pressing pressures of 1. 5, 3. 0, 4. 5 and


6. 0 t/ cm 2
2
,, respectively.

Table 2. 10 Magnetic properties of polymer bonded anisotropic NdFeB magnets made by hot
compression molding.
Pressing
Sample Pressure B, Hei
He; He ( BH)max
C BH)ma,
Polymer temp.
name
Ct/cm 3 ) CC) (mT)
CmT) (kA!m)
CkA!m) (kA!m)
CkA!m) CkJ/m 3 )
(kJ/m

Composition A

NOI
N01 Nylon 12 1.5 190 633 860 395 66

N02 Nylon 3.0 190 695 850 422 78

N03 Nylon 45
4.5 190 758 853 451 92

N04 Nylon 6.0


60 190 842 839 487 113

Composition B

POl PPS 1.5 290 645 847 400 68

P02 PPS 3.0 290 682 828 410 75

P03 PPS 4.5 290 804 801 461 102

P04 PPS 6.0


60 290 841 785 473 124

C4) Compression molded SmCo magnets


Raw materials with the composition SmCCoO 625 FeO28CuO.07ZrO
O. 625 Feo.2s Cuo 07Zr0025)S were
025)844were

melted in a induction melting furnace. The melted liquid alloy was then poured
into a copper mold at a predetermined speed to produce an ingot. About 90% 90 %
volume percent of the desirable columnar grains within the ingot was obtained
by adjusting the liquid alloy temperature, the speed of pouring liquid alloy into
the copper mold, the cool ing rate of the ingot, etc. The ingot was then
cooling
solution treated at 1140 to 1200 'c for 2 to 10 h, and then heat-treated at 750
to 850 'c for 5 to 20 h followed by slow cooling to 400 'c at a rate of 1 to 1.5 1. 5
'c /min. The above ingot was then crushed under the protection of argon
atmosphere, followed by milling to achieve the desired particle size and
distribution. The powder was then coated by approximately 3. Owt. % epoxy.
The above coated powder was then compression molded at various
temperatures under a pressure of 8 tlcm t/ cm 2
2
.. The maximum energy product,
CBH)max,
CBH) max. versus pressing temperature is shown in Fig. 2.14. 2. 14. An increase in
BH) max of 11 % was achieved by increasing pressing temperature from room
CBH)max
temperature to 180 'c .
52 Jinfang Liu and Michael Walmer

140

120
~ ~
~
-
~

E L.r
U"

2
~ 100
g~
~ 80

60

40 80 120 160 200


(OC )
Pressing temperature (C

Figure 2. 14 The maximum energy product, (BH) max'


rna>' versus pressing temperature

for SmCo bonded magnets compression-molded under a pressure of 8 t/ cm 2 .


t/cm

It is concluded that hot compression molding can increase the density and
magnetic properties of bonded rare-earth magnets. But hot compression
molding process is a bit more complex than conventional compression
molding. No high-volume production using this process has been reported yet.

2.3.3 Bonded Magnets Produced by Injection Molding Process

The injection molding process has the advantage of producing components with
complex shapes with consistent accuracy. Figure 2. 15 is a sketch of an
injection-molding machine. The rare-earth alloy powders, such as isotropic
NdFeB and anisotropic SmCo powders, are blended with thermoplastic
materials, such as polyphenylene sulfide (PPS) and nylon. Polymer additives,
such as antioxidants, plasticizers, lubricants, etc. , are also needed in most
cases. The blended mixtures are then intensively mixed together by a
compounder to form a compound. The compound is then chopped into small
granules and fed into the injection-molding machine, which can be a horizontal
or vertical type. In the case of multi-component injection molding, a
subassembly consisting of a bonded magnet and other functional elements,
such as gears and shafts, can be made, eliminating or minimizing the
assembly cost. Also, the inter-parts tolerances are not of concern because the
subassembly is made by one set of tooling. In the case of insert molding, the
component is first placed in the molding tool, the tool closed and the magnet
material injection molded around the component. As the polymer bonded
magnet material shrinks during cooling, the component is held tightly in place.
One of the greatest advantages of the injection molding process is that a high
degree of shape complexity is possible. The disadvantage of injection molding
is the high tooling cost (Paju, 2000). Table 2. 11 shows the typical magnetic
Process and Magnetic Properties of Rare-Earth Bonded Magnets 53

properties of the injection-molded magnets.

Table 2.11
2. 11 CBH )mex
Typical maximum energy product (BH) me< of injection molded rare-earth
magnets.
magnets

BH)ma,
(CBH)me<

.~_T_y_p_eS_o_f_in_ie._c_ti_o_n_m_o_ld_e_d_m_a_g_ne_t_s__---tf-- ---_-_-_-(_~_~_O-;)-------T------(-~~/~;-)-----
Types of injection molded magnets
__ CMGOe) CkJ/m 3 )

Isotropic NdFeB bonded magnet 5-6


5- 6 I 40-
40 - 48
------------~----------------- ----- +--
----- ------------
~._-----_._--~ _ .. - - ~ - - ---
_ - - ~ - - - - -

Anisotropic NdFeB bonded magnet 8-


8 - 11 72- 88
~---------------+----------_ ..-
Anisotropic SmCo bonded magnet 8.5-9.5 68-76

Fi~ure 2. 15
I<igure t5 A sketch showing an injection molding machine.

2.3.4 Bonded Magnets Produced by Extrusion Process

The extrusion process is commonly used for flexible bonded ferrite magnets.
The advantage of this process is continuous production of magnets of two-
dimensional geometries, such as strips or sheets of virtually unlimited length.
The strips or sheets of magnet material can then be punched into smaller parts
with simple geometries, such as discs and washers. Recently, the extrusion
process has also been used for the production of bonded rare earth magnets.
Figure 2. 16 is a sketch showing a horizontal screw extruder. The main
components of a screw extruder include the feed hopper and throat, extruder
screw, extruder barrel, the die, drive system, heating and cooling
cool ing systems,
and the control system. Ram extruders are also used for rare earth bonded
magnets. Figure 2. 17 is a sketch of a vertical ram extruder (Akioka, 1998).
High pressure can be developed during ram extrusion, which is advantageous
for the production of highly loaded bonded magnets.
54 Jinfang Liu and Michael Walmer

- Feed hopper

Screw

Figure
I<'igure 2.16
2. 16 A sketch showing a horizontal screw extruder.
extruder,

Hydraulic
cylinder

Hydraulic drive

Frame
Connected to the
::::.--------- feed hopper
::::-------
~:;;:Z:ZZ2Z1
~~:zzJ

Heater

Throat ~m--fmm.:l
"",,,*+-=1, Thermal insulation portion

---",'/
Die - --'//
~f"'--"'T~
~,........,.~--
_ _ Cooling apparatus

Figure 2. 17 A sketch of a vertical ram extruder.


extruder,
Process and Magnetic Properties of Rare-Earth Bonded Magnets 55

In extrusion, there are many parameters that can affect the quality of the
magnets. Extrusion temperature, the difference of temperature settings in
various zones, ram extrusion pressure, slow ram speed, type of
thermoplastics, magnet material loading, design of the tooling, and cooling
techniques are all important parameters for the ram extrusion process.
Automated feeding system, cutting devices, and conveyors can be added for
large volume production line.
Figure 2. 18 shows a comparison of demagnetization curves of bonded
NdFeB isotropic magnets produced by compression, extrusion and injection
molding processes.
800 ,---~---,-~~-.-~~-,---~--,c--~--,
,-------,---~---,------,c__----,

600 1--~--+~~-I-~~+-,~,.L::.-f---=",,"'9I
1-----+---I----+---"~"""""-I__-=-"oI';lI

E
~ 400 1--~-I-~~--V~~-l+--,~f,L-j'-l----I
I-----l------V~-_HL--,-I+-J'-I--__I
...
...,o
'" 200 I-----+--~-I--+----_++_-+-__/__II__-__I
I--~--+~i---I-~-I-+l---+---I-Jf--~--I

o -800 ---fi00
-600 -400 -200
Field H(kOe)

Figure 2. 18 Demagnetization curves of bonded NdFeB magnets produced by


compression, extrusion, and injection molding, respectively.
respectively_

2. 4
2.4 Polymers and Polymer Additives for Bonded Rare-
Earth Magnets

A variety of binders can be used for bonded magnets, including thermoplastic


and thermosetting resins and elastomers. Table 2. 12 shows some of the
characteristics of polymers for bonded rare-earth magnets. Ikuma,
lkuma, et al.
claimed that the flow ability for a bonded magnet composition containing at
least two types of thermoplastic resins (each with melting points within 50 'C)
is better than a composition containing only one type of thermoplastic resin
(Ikuma et al. , 1999).
56 Jinfang Liu and Michael Walmer

Table 2. 12 Typical characteristics of polymers used for bonded rare-earth magnets.


(</>3. 175 mm
Water absorption (<1>3.175
Melting point Density thick specimen)
Polymer (% )
('C)
CC) (g/cm 3 )
24 h Saturation

Epoxy Thermoset 11-11.40


1. 11- .40 0.08- 0.15
0.08-0.15

Nylon 12 160 - 209


160-209 01 - 11.02
11.01- 02 0.25-
0.25-00.30
30 0.75-1.6
0.75 - 1.6

Nylon 6/12 195 - 219


195-219 1.06-1. 10 0.37 - 1.0
037-1.0 25-3.0
2 5- 3 0

Nylon 6 210 - 220 1.


1. 12-1. 14 1.3- 1.9
1.3-1.9 8.5- 10
8.5-10

PPS 285- 290 1. 35


1.35 >0.02-0.07
>0.02-007

The addition of an antioxidant compound was found to be one of the most


important polymer additives leading to improvements in the compounding and
extrusion processes. The sol id loading level can be increased as much as 2
solid 2%%
wt by the addition of proper antioxidant.
Various levels of lubricants have also been used in the ram extrusion
experiments. Table 2. 13 shows the correlation between lubricants and the
extrusion process. It can be seen that the percentage of lubricant needs to be
optimized in order to have a smooth extrusion process.

Table 2. 13 Effect of lubricants on the ram extrusion process.


Extrusion
MQP-B
MOP-S Nylon 12 Antioxidant Lubricant
temperature Comment
(wi
(wI. %) (wt.
(wI. %) (Wt. %)
(WI (wi %)
(wI.
('C)
CC)

94 4.3 0.5
05 1.2 230 Difficult to extrude

94 4.6
4 6 0.5 0.9 230 Runs smoothly

94 4.8 0.5
05 0.7
0 7 230 Runs smoothly

94 5.0 0.5 0.5


05 230 Difficult to extrude

92.5 5.8
58 0.5 1.2 230 Runs smoothly

92.5 6.3 0.5 07


0.7 230 Runs smoothly

92.5 6.5 0.5 0.5 230 Runs smoothly

92.5 6.8 0.5 02


0.2 230 Runs smoothly

92.5 7.0 0.5 0 230 Difficult to extrude

An amine based titanate coupling agent, L1CA44 , was used in the


compression molding experiments. The chemical name of L1CA44 is neopentyl
(diallyl) oxy, tri (N-ethylene-diamino)
( N-ethylene-diamino) ethyl titanate. MQLP-B powder was
coated using the coupling agent L1CA44. The coated powder was then
Process and Magnetic Properties of Rare-Earth Bonded Magnets 57

compression molded at room temperature. The test results are shown in


Table 2. 14. It can be seen that L1CA44 coupl ing agent can improve the density
and the magnetic properties. Chen et al. reported that L1CA44 could also
improve the corrosion resistance of the bonded magnets (Chen et al.,
al. , 1998).

Table 2. 14 Effect of coupling agents on the density and magnetic properties of epoxy
bonded magnets.
Compression
Compression MQP-B Zinc
MQP-S Epoxy L1CA44 Density (BH)ma,
Pressure Stearate

(wI. %)
(wt.%) (wt. %)
(wt.%) (wI. %)
(wt.%) (wI. %)
(wt.%) (g/cm 3 ) (MGOe) (kJ/m 3 )

2 98 1.9
19 o 0.1 5.65 9.71
971 77.3
77. 3

2 98 1.5 0.4 0.1 5.74 9.80


980 78.0
4 98 1.9
.9 o 0.1 5.80 9.82
982 78.2
4 98 1.55 0.4 O. 5.84 9.90 788
6 98 1.99 o o
O. 5.85 9.91 78.9
6 98 .5 0.4 o 5.90 9.97 79.4
8 98 1.99 o O. 5.90 9.98
998 79.4
8 98 1.5
.5 0.4 0.1
O. 5.99 10.1 80.4
10 98 .9
1.9 o 0.1
O. 5.98 10.01 79.7
10 98 1.5
.5 0.4 0.1
O. 6.09 10.4 82.8

2. 5 Environmental Stability of Polymer Bonded Magnets

Some of the polymer-bonded magnets are used at elevated temperatures and/


or other severe environmental conditions. It is therefore important to
understand the effects of the environment on the magnetic properties of bonded
magnets. Figure 2. 19 shows the total flux loss of epoxy-bonded magnets at
various temperatures as a function of time (Magnequench, 2000a). The
geometry has a significant impact on the total flux loss due to the differences of
self-demagnetization factor. Careful consideration of the shape and size of the
magnets during the design stage would help to increase the maximum possible
operating temperature.
Surface epoxy coating for NdFeB bonded magnets can significantly
improve the environmental stability. Samples Ax and Bx (x = = 1, 2, 3 and 4)
were used to test the long-term stability: Samples AI' A 2z ,, A3 and A4 are
bonded NdFeB isotropic magnets with epoxy surface coating, while samples
58 Jinfang Liu and Michael Walmer

o 200 400600 800


Time (h)
Figure 2. 19 Total flux loss versus time for compression molded NdFeB magnets at
various temperatures (top: B/H= -5, middle: B/H= -3, and bottom B/H= -2).

8 1 , 8 2 , 8 3 and 8 4 are bonded NdFe8 isotropic magnets without surface


coating. The dimensions for samples Al , A 2 , A3 , A44 , 8 11 , 8 2 , 8 3 and 8 4 are
the same. The outer diameter (00), inner diameter (10), and thickness
(THK) of these samples are 34. 8, 25, and 31. 25 mm, respectively. All
above samples were magnetized through diameter. Flux and weight changes
for samples Al and 8 1 were measured following the procedures described in
Fig. 2.20.

(I) Magnetize the samples after weighing them

oven at T 1 far
(3) Heat treat in aven for 60 min

(4) Remove
Remave the samples from
fram the oven caal ta
aven and cool to roam
room temperature

(5) Test the weight and the flux afthe


of the samples

(8) Heat treat in aven +25) C far


oven at (T 1j +25)'C for 60 min

(9) Repeat 4 to (T 1 +25)=300 'c


ta 8 until (T[+25)=300 C

Figure 2. 20 Experimental procedures for the measurement of the flux and weight
changes versus temperature.
Process and Magnetic Properties of Rare-Earth Bonded Magnets 59

Figure 2.21 shows the total flux loss versus temperature for both coated
and uncoated samples. Below 220 C , the total flux loss changes gradually for
both samples. When the temperature is higher than 220 C , the total flux loss
of the non-coated sample (B,)
(B l ) increases very quickly, while that of the coated
sample (AI) keeps changing gradually. This indicates that oxidation starts at
about 220 C for the non-coated samples. This is confirmed by the weight
change measurement as shown in Fig. 2. 22. Below 220 C , the flux loss can
be recovered by re-magnetizing the samples at room temperature. But above
220 C , the flux loss can be only partially recovered due to oxidation as shown
in Fig. 2.23.
20
SampleA
SampleA]I
o (Epoxy coated)

[~ -20 ~~
"'"'enen
52
.-'? -40 \
~
"
::J

;;: -60
-60 S,mp"",
S=pl,3, "~
,~
) -
'El
~
r=: -80
(No furface
surface 1 coatin
1 1 r
coating)
I

-100
\
o 50 100 150 200 250 300
Temperature CC )

Figure 2. 21 Total flux loss versus temperature for bonded NdFeB magnets magnetized
through thickness (Dimension: OD 34.8 mm, ID 25 mm, THK 31 . 25 mm).
mm) .

2.0
I I
~
~

C <l)
Q)
OJ)
b.O
1.5 Sample B
(Nf
1
B]I
coatin g
(N surface coating )1J
1.0
"u'"
/
oj
.:=
..<:

.E
1:
II
OJ)
b.O 0.5 Sample AlA]
~
~
(With epoxy coating)
J .1 J.I~_
~
o ~

1
o 50 100 150 200 250 300
Temperature CC
('C ))

Figure 2. 22 Weight change versus temperature for bonded NdFeB magnets magnetized
through thickness (Dimension: OD 34.8 mm, mm. THK 31 ..25
mm. ID 25 mm, 25 mm) .

Both samples A2 and B2 were stabilized at 120 C for one hour before a
long-term aging test at 100 C for 4000 h. The test result is shown in
Fig. 2.24.
2. 24. The flux decrease for the surface coated sample is insignificant
after 4000 h at 100 C. It should be pointed out that the total flux loss at
elevated temperatures is dependant upon the size and geometry of the
60 Jinfang Liu and Michael Walmer

~ o0 "I1 l1 ~.-7
~
~
~-IO
-10
SampleA
Sample AI1 ~
-----
.3
..2
~ -20
(With epoxy coating)
+-f-- \ f--
1 1
r'\
r~
,
<;::
<;:::
Sample 8B 1I ~
"
<l)
----
Z -30
:0
(No surface coating)
'
.~

~ -40
\
~
~
..::~
-50
o 50 100 1150'
50' 200 250 300
Temperature (C
(OC )

Figure 2. 23 Irreversible flux loss versus temperature for bonded NdFeB


magnets, magnetized through thickness (axially) (Dimension: 00 34.8
34. 8 mm,
1025 mm, THK 31.25 mm).

samples. Irreversible losses due to oxidation or corrosion will be larger for


smaller samples due to their larger S / V ratio, where S is the total surface
area and V is the volume of the sample.
Anti-corrosion evaluations for the surface coated samples A 3 , A 4 and un-
coated samples 8 3 , 8 4 were conducted in a 100 % humidity environment at
room temperature. Samples A3 , A4 , 8 3 and 8 4 were put into sealed containers
where the relative humidity was 100 % .
o 000
~:O Po
00

~o
DO 0
C\ 00
Q
DO 0
A2
0
~ -{).5 --u
0 J=-o--o-o---
'U

~ a0
c:
~
oj
-1.0 a0
x::l
0
4: 0
0
'Ei -1.5 e-----
f...-.-.----I-. ----- f----------- ---
~-

~ 8B22 0
ao 0
0
000
000
-2.0
o 1000 2000 3000 4000
Time at 100C
100C (h)

Figure 2.24 100C


2. 24 Long-term stability of bonded magnets at 100 C (Dimensions:
0034.8 mm, 1025 mm, THK 31.25 mm).

The Flux and weight changes versus time (up to 8000 h) under 100 %
relative humidity were measured regularly. The experimental results are
shown in Figs. 2. 25 and 2. 26, respectively. The non-coated sample 8 3 was
rusty after only two days while the coated sample A3 showed no rust at all.
After 8000 h, the non-coated sample 8 3 was very rusty with a significant
weight increase, while the weight of the coated sample A3 remained virtually
Process and Magnetic Properties of Rare-Earth Bonded Magnets 61

unchanged. Again, the percentage of the total flux loss is related to the
geometry of the samples. Smaller samples will have proportionally larger
irreversible flux loss because of the larger surface to volume ratio (corrosion
starts on the surface of the samples). Both the sample weight and flux level
did not change significantly for the coated samples A3 and A 4 . There appeared
to be no rust at all for the surface-coated sample A3 after 8000 h under 100 %
relative humidity.

I
o v 0 w
~
~ W
~ u. 0 0
0
w
1u
-1
-I
\SampleA 31
\Samp1eA _ -
0
~
e (Epoxy coated)
0
~-2
<I>

.2
B 0
_SampleB} ____
~ -3 I--- Sample B 1 ---
c;::
<;::
0
(No coating)
'3 0
~ -4
0
-5

o 2000 4000 6000 8000


Time in 100% relative humidity environment (h)

Figure 2.25
2. 25 Total flux loss of bonded NdFeB magnets in 100
100%% relative humidity
environment (Dimension: 00 34.8 mm, 10 25 mm, THK 31.25
31 .25 mm) .

Ell B3(No coating)


2.0
o B4(No coating)
~ 1.5
oa A3(Coating)
JIll A4(Coating)
lIIl
~
c
.g 1.0
1:
.~ 0.5
~
0.0
o 1000 2000 3000 4000 5000
Time in 100%
J 00% relative humidity environment (h)

Figure 2. 26 Weight change versus time for bonded NdFeB magnets in 100 % relative
humidity environment (Dimension: 00 34.8 mm, 10 25 mm, THK 31 . 25 mm) .

2.6 Summary

The bonded magnet market is expected to grow as the consumer electronics


industry grows. Careful selection of bonded magnets in various applications is
62 Jinfang Liu and Michael Walmer

needed to optimize the performance and reduce the cost. Improvements of


magnetic properties of bonded magnet products will help in the effort to find
new applications
appl ications and niche markets.
New developments in processes and magnet powders are two of the keys
2.15
to improvements of magnet performance. Table 2. 15 summarizes the process
and magnetic properties of bonded magnets (Ormerod and Constantinides,
1997). Some of the processes and magnet powders listed in Table 2. 15 are
not covered in this chapter.

Table 2. 15
IS Summary of the processes and magnetic properties of bonded magnets.
Process Compression molding Extrusion Injection molding
Iniection Calendering
Thermoplastic or
Binder Thermoset Thermoplastic Elastomer
elastomer
eopxy, acrylic.
Thermoset: eopxy. acryl ic, phenol ic;
Thermoplastic: polyamides (such as nylon 6.6/6,6/12).
6, 6/6, 6/12), polyester, polyphenylene sulfide.
sulfide,
Examples of binders
polyvinyl chloride, low density polyethylene;
Elastomer: nitrile rubber.
rubber, vinvl
Flexibility of the
Rigid Rigid or flexible Rigid Flexible
end product
Maximum energy product (BH)ma,
(BH)max
Magnetic powders
(kJ/m 3 ) (MGOe) (kJ/m 3 ) (MGOe) (kJ/m 3 ) (MGOe) (kJ/m 3 ) (MGOe)
Sm(CoCuFeZr),
104 -135 13-17 48-80
48 - 80 6-10 48 - 80 6-10 N/A N/A
(anisotropic)
SmCos
SmC05 (anisotropic) 64-96 8-12 32-80
32- 80 4-10
4 -10 32-72
32 -72 4-9 N/A N/A
NdFeB(anisotropic) 111-127 14-16
14 -16 N/A N/A N/A N/A N/A N/A
NdFeB( isotropic) 72-80
72- 80 9-10 32-64 4-8 32-48 4-6 24-40 3-5
Ferrite( isotropic or
N/A N/A 32-119 04-15
0.4-1.5 40-135 0.5-1.7 4.0-12.7
4.0-135 0.5-17 4 0- 12 7 05-16
05-1.6
anisotropic)
SmFeN 64-119 8- 15
8-15 N/A N/A N/A N/A N/A N/A

Compared to bonded NdFeB magnets, bonded SmCo magnets have better


high temperature capability. Figure 2. 27 shows the comparison of high
temperature performance between bonded SmCo and NdFeB magnets. Both
SmCo and NdFeB bonded magnet samples used in this experiment are
cylindrical shape, 10 mm in diameter and 10 mm long. The samples were
magnetized along the length. Polyphenylene sulfide was used as a binder for
the bonded SmCo magnet samples, while high temperature epoxy was used for
bonded NdFeB magnet samples. It can be seen that the total flux loss of SmCo
bonded magnets is very small compared to that of bonded NdFeB magnets at
elevated temperatures. Therefore, bonded SmCo magnets are highly
recommended for applications at elevated temperatures.
Process and Magnetic Properties of Rare-Earth Bonded Magnets 63

SmCo
o
~
~ -20
~
~
.3
x -40
\
NdF'~
Ndr,~
:::l
<;:
3 e-
>--
?=: -60

-80
o 100 200 300
Tem perature CC
Temperature )
Figure 2. 27 Comparison of high temperature stability between bonded
SmCo and NdFeB magnets (Dimension: 10 mm x 1Omm)
10mm)..

Axial magnetization Diametrical magnetization

Multi-polar magnetization
Radial magnetization
on both ends

Multi-polar magnetization Multi-polar magnetization


on the circumference on one end of a disc
Figure 2.28 Examples of magnetization patterns of bonded NdFeB isotropic magnets.

If the maximum operating temperature is below 150C, preferably below


c , bonded NdFeB isotropic magnets can be considered. Isotropic bonded
120 C,
64 Jinfang Liu and Michael Walmer

magnets can be magnetized into various patterns, including axial, diametrical,


radial and multi-polar magnetization. Figure 2. 28 gives a few examples of
magnetization patterns that are possible with bonded isotropic magnets. Multi-
poles on the outer or inner diameters of a sleeve-shaped bonded isotropic
magnet are among the popular choices for small motor applications.
Figure 2. 29 shows the trapezoidal and sinusoidal flux configurations on the
outer diameter( 00)
OD) of a 4-pole bonded NdFeB magnet with outer diameter of
21.5
21 . 5 mm, inner diameter(
diameter(IO)
ID) of 15.4 mm, and thickness of 6.
6.11 mm. The flux
configuration can be changed by the design of magnetizing coils and/or the
variation of magnetizing parameters. Figure 2.30 is a photo showing extruded
rare earth magnets produced at Electron Energy Corporation (Electron,
2000b). Figures 2.31
2. 31 and 2.32 are photographs showing compression molded
ring and disc magnets, respectively (Electron, 2001).

Trapezoidal 4-pole magnetized on the 00


OD

-+-----\
+ 11
------- '-
'-,-----\-----,360

Angle (')

Sinusoidal 4-pole magnetized on the 00


OD

-+------'<------,<------"r-----.360
---Jl---------'<---------t'------"r-------,360

Angle (')

Figure 2. 29 Trapezoidal and sinusoidal magnetization of a 4-pole bonded NdFeB


magnet (magnet dimension: 0021.5
OD 21.5 mm, 10
ID 15.4 mm, Thickness 6. 1 mm).

Figure 2.30 Photograph showing bonded rare-earth magnets produced by


extrusion process.
Process and Magnetic Properties of Rare-Earth Bonded Magnets 65

000000
o 000
"s
000 000 <;

0
o00000000 0
000000000000
0000000000

Figure 2.31 Photograph showing bonded ring magnets produced by


compression molding .





\ ..
000

'\
Figure 2. 32 Photograph showing bonded disc and ring magnets
produced by compression molding.

References
Ahmed, F. M., A. Ataie, A. J. Williams, and I. R. Harris. In: Proceedings
of the 6th European Magnetic Materials and Applications Conference.
Vienna, Austria, Sept. 4 - 8( 1995)
Ak ioka, K. European Patent. EP 0865051 A 1( 1998)
Akioka,
Bradbury, S. Powder Metallurgy Equipment Manual, 3rd ed. ed. By Powder
Metallurgy Equipment Association, Division of Metal Powder Industries
USA( 1986)
Federation, New Jersey, USA(1986)
Chen, C. H., M. S. Walmer, M. H. Walmer, S. Liu, E. Kuhl and G.
Simon. J. Appl. Phys. 83: 6706 (1998)
Chen, C. H. , Marlin
Marl in S. Walmer, Michael Walmer, Jinfang Liu Sam Liu and
G. E. Kuhl. J. Appl. Phys. 87: 6719 (2000)
Chen, Q. , J. Asuncion, J. Landi, B. M. Ma. In: FF-ll,
FF-ll , 43rd Annual Conf.
On MMM, Miami, Florida, 11: 9-12(1998)
Coehoorn, R., D. B. de Mooij, J. P. W. B. Duchateau and K. H. J.
Buschow. J. Phys. (France) C 8: 669 (1988)
66 Jinfang Liu and Michael Walmer

Coehoorn, R. , D. B. de Mooij and D. de Waard. J. Magn. Magn. Mater.


80: 101 (1989)
Croat, J. J. , J. F. Herbst, R. W. Lee and F. E. Pinkerton. Appl. Phys.
Lett. 44: 148( 1984)
Croat, J. J. J. Mater. Eng. 10: 7 (1988)
Croat, J. J. In: 11th
11 th Int. Workshop on RE Magnets and Their Applications.
Pittsburgh, PA, USA, 10: 7(1990)
Croat, J. J. In: 13th Int. Workshop on RE Magnets and Their Applications.
Birmingham, UK, 9: p.65(1994)
Croat, J. J. J. Appl. Phys. 81: 4804 (1997>
Dahlgren, M., X. C. Kou, R. Grossinger, J. F. Liu, I. Ahmad, H. A.
Davies and K. Yamada. IEEE Trans. Magn. 33: 2366 (1997a)
Dahlgren M. , R. Grossinger, E. de Morais, S. Gama, G. Mendoza, J. F.
Liu and H. A. Davies. IEEE Trans. Magn. 33: 3895 (1997b)
Dahlgren, M. , R. Grossinger, J. F. Liu and H. A. Davies. In: L. Schultz and
K. - H. Muller eds. Rare Earth Magnets and Their Applications, 1: 253
(Proc. 15th IntI.
Inti. Workshop on RE Magnets and Their Applications, Aug.
30 to Sept. 3, 1998, Dresden, Germany)
Davies, H. A. Of the 9th IntI. Symp.
A.,, J. F. Liu and G. Mendoza. In: Proc. Ofthe9thlntl.
On Magnetic Anisotropy and Coercivity in RE-TMRE- TM Alloys, Volume 2. Sao
Paulo, Brazil, p. 251(1996)
Electron Energy Corporation. Polymer Bonded Magnets Catalog. Landisville,
Pennsylvania, USA, 2000a
Electron Energy Corporation's website: www.electronenergy.com , 2000b
Electron Energy Corporation. Internal Technical Report, 2001
GolI, D., M. Seeger, H. Kronmuller. J. Magn. Magn. Mater. 185: 49
(1998)
( 1998)
Hadjipanayis, G. C. and W. Gong. J. Appl. Phys. 64: 5559 (1988)
Hadjipanayis, G. C. and R. F. Krause. IEEE Trans. On Magn. 31: 3596
( 1995)
Herbst, J. F. In: G. J. Long and F. Granddjean eds. Supermagnets, Hard
Magnetic Materials. Kulwer Acad. Pub. p. 69 69(( 1991)
Herchenroeder, J. Magnequench Quality
Qual ity Document 1 (2000)
Ikuma K., T. Ishibashi and K. Akioka. United State Patent 5,888,416, 5, 888, 416,
(1999)
( 1999)
Kim, A. S. J. Appl. Phys. 81: 5609 (1997)
Kneller, E. F. and R. Hawig. IEEE Trans. Magn. 27: 3588 (1991)
Liu, J. F.,
F. , H. A. Davies and R. A. Buckley. In: Proc. 13th Inti. Workshop
on RE Magnets and Their Applications. Sept. 11. Birmingham, U. K., K. , p.
79( 1994a)
Liu, J. F.,
F. , I Ahmad, H. A. Davies and R. A. Buckley. In: Proc. 8th Inti.
Symposium on Magnetic Anisotropy and Coercivity in RE- TM Alloys. Sept.
15. Birmingham, U. K., p. 169(1994b)
Liu, J. F., I Ahmad, H. A. Davies, P. Z. Zhang, S. G. Huo and R. A.
Process and Magnetic Properties of Rare-Earth Bonded Magnets 67

Buckley. In: Proc.


Proc . 8th Inti. Symposium on Magnetic Anisotropy and
Coercivity
CoercivityinRE-TMAlloys.
in RE-TM Alloys. Sept. 15, Birmingham, U. K., p. p.161
161 (1994c)
Liu, J. F. and H. A. Davies. J. Magn. Magn. Mater. 157/158: 29 (1996)
Liu, J. F. , T. Chui, D. Dimitrov and G. C. Hadjipanayis. Appl. Phys. Lett.
73: 3007 (1998a)
Liu, J. F. , Y. Zhang, Y. Ding and G. Hadjipanayis. In: Rare Earth Magnets
and Their Applications, edited by L. Schultz and K. - H. Muller, volume 2,
p.607 (Proc. 15th Inti. Workshop on RE Magnets and Their Applications,
Aug. 30 to Sept. 3. Dresden, Germany ( 1998b)
Liu, J. F. Progress Report on Bonded HDDR Magnets. Electron Energy
Corporation ,_andisville , Pennsylvania, USA, March (1999a)
Liu, J. F., 't. Zhang, Y. Ding, D. Dimitrov and G. C. Hadjipanayis. J.
Appl. Phys. 85: 5660 (1999b)
Liu, J. F.,
F. , Y. Zhang, D. Dimitrov and G. C. Hadjipanayis. J. Appl. Phys.
85: 2800 (1999c)
Liu, J. F., Y. Ding, D. Dimitrov and G. C. Hadjipanayis. J. Appl. Phys.
85: 1670 (1999d)
Liu, J. F. , Y. Zhang and G. Hadjipanayis. J. Magn. Magn. Mater. 202: 69
(199ge)
Liu, J. F. and G. Hadjipanayis. J. Magn. Magn. Mater. 195: 620 (1999)
Magnequench International. NdFeB Product Catalog. Indianan, USA, (2000a)
Magnequench International. MOl MQl Application Bulletin. Indianan, USA,
(2000b)
Manaf, A., R. A. Buckley, H. A. Davies and M. Leonowicz. J. Magn.
Magn. Mater. 101: 360 (1991)
Matzinger, M. , J. Fidler, A. Fujita and I. R. Harris. In: Proceedings of the
6th European Magnetic Materials and Applications Conference. Vienna,
Austria, Sept. 4 - 8( 1995)
Nakayama, R. and T. Takeshita. J. Appl. Phys. 74: 2719 (1993)
Nakayama, R. R.,, T. Takeshita, M. Itakura, N. Kuwano and K. Oki. J. Appl.
Phys. 70: 3770 (1991)
Nesbitt, E. A., R. H. Willens, R. C. Sherwood, F. Buehler and J. H.
Wernick. Appl. Phys. Lett. 12: 361(1968)
361( 1968)
Ojima, T. , T. Tomizawa, T. Yoneyama and T. Hori. IEEE Trans. Magn.
MAG-13: 1317 (1977)
Okonogi, I. and T. Shimoda. US Patent 4,536,233, August 20 ( 1985)
Ormerod, J. and S. Constantinides. J. Appl. Phys. 81: 4816 (1997)
Paju, M. Multi-component injection molding for bonded magnet assemblies.
assembl ies.
In: Intertech Polymer Bonded Magnets 2000, March 27 - 29. Nashville,
TN, USA(2000)
Perry, A. J. and A. Menth. IEEE Trans. Magn. MAG-ll: 1423( 1975)
Ray, A. E. and K. J. Strnat. IEEE Trans. Magn. MAG-8: 861 (1972)
Ray, A. E. J. Appl. Phys. 55: 2094 (1984)
Ring, J. Dynamics of bonded NdFeB permanent magnet market. In: Intertech
68 Jinfang Liu and Michael Walmer

Polymer Bonded Magnets 2000, March 27 - 29. Nashville,


Nashvi lie, TN, USA
(2000)
Takeshita, T. and K. Morimoto. J. Appl. Phys. 79: 5040 (1996)
Takeshita, T. and R. Nakayama. In: Proceedings of the 10th International
Workshop on Rare-Earth Magnets and Their Applications, Vol. 1. Kyoto,
Japan, p.551(1989)
Uehara, M. , H. Tomizawa, H. Hirosawa, T. Tomida and Y. Maehara. IEEE
Trans. Magn. MAG-29: 2770 (1993)
Wohlfarth, E. P. Ferromagnetic Materials, Volume 3. North-Holland
Publ ishing Company, Amsterdam, 1982
PubIishing
Zhang, Y. , M. Corte-Real, G. Hadjipanayis, J. F. Liu, M. Walmer and K.
M. Krishnan. J. Appl. Phys. 87: 6722 (2000)
3 Laser Processing of Magnetic Materials

Yi Liu, Min Zheng, D. J. Sellmyer and J. Mazumder

3. 1 Introduction

The fascinating laser beams are ideal tools for materials processing. The word
laser is an acronym for light amplification by stimulated emission of radiation.
Laser is the most convenient way to deliver pure energy into any material with
the option of when, where, and how intense. Naturally, a laser beam provides
a unique means for materials processing.
Various lasers have been widely used for materials processing (for a
review see Ready, 2001). Following are some important examples: CD Laser
surface treatment by phase transformation resulting from rapid heating and
rapid cooling rate generated by laser radiation: This includes laser surface
melting, laser surface hardening/treatment, laser surface alloying, and laser
al., 1994, 1995); (2)
cladding (Liu et ai., ~ Laser direct metal deposition
(Mazumder et al., 1997); @ Physical vapor deposition by laser ablation
laser patterning / laser lithography and @ Laser machining such as laser
cutting, laser drilling, and laser welding. Materials processed by laser
techniques cover a wide variety including steels, high temperature alloys,
intermetallic
intermetall ic compounds, semiconductors, super-conducting materials, and
magnetic materials.
However, laser processing of magnetic materials is quite limited
compared to mature applications such as laser drilling, laser welding, and
laser surface modification. In this chapter our purpose is to give an introduction
to laser materials processing to colleagues in magnetic materials research,
and address the great potential of using lasers for magnetic materials
processing. We first present the novel characteristics of various lasers and the
principle of laser materials processing. We will review our recent results of
grain size and texture control by laser induced phase transformation, nanodot
array fabrication by interferometric laser lithography, and other possible
applications. Issues associated with laser materials processing such as laser
materials interaction, temperature field, mass transfer, phase transformation,
solidification process, and microstructure-properties relation will be
discussed.
70 Vi Liu et al.

3.2
3. 2 Laser Characteristics and Laser Facilities for Materials
Processing

3.2.
3. 2. 1 Laser Characteristics

Laser materials processing involves interdisciplinary science and engineering.


First, the selection of a proper laser for a specific material synthesis involves
the wavelength, the power density and the mode. Second, a proper optical
deliver system is required. Third, a specimen handling system, such as a
multiple-axes stage, or a vacuum chamber is needed. The most important
characteristics of lasers for materials processing are: wavelength, operation
mode (continuous or pulsed), power density, monochromaticity, coherence,
and beam mode (beam intensity profile). Table 3. 1 gives various lasers
available for materials processing.

Table 3.1
3. 1 Lasers and their characteristics.

Laser Wavelength Power range Mode


CO22 ~m
10.6 IJm Up to 50 kW P&CW()
P&CW'"
Solid state lasers o 66-2.94
0.66-2. 94 ~m
IJm Up to 3.5
3 5 kW P&CW
Ion lasers Visible 275 - 1092 nm A few walts
watts CW
Excimer 158 - 351 nm
158-351 40-100 W P
X-ray lasers 1-
1- 47 nm -12
- 12 MW(2)
MW<2l P
(1) P represents pulsed operation mode and CW represents continuous wave mode.
(2) Averaged within the pulse duration.

Laser wavelength has two major effects: First, laser wavelength


determines the spatial resolution achievable. For example, both the spot size
of the beam and the smallest pattern in laser lithography are limited to the
same scale of the wavelength. Second, laser wavelength affects absorbability
and absorption depth. Different materials can show substantial difference in
absorbing
absorb ing laser energy. The absorption depth is proportional to wavelength.
Laser mode refers to the operation mode: Either a continuous wave or a
pulsed mode. Pulsed operation is occasionally used for reduction of the heating
of the laser. A Q-switch can be used to concentrate the laser energy into a
pulse and mode locking can shorten the duration of the pulse. The intensity of
the pulsed laser can be several orders higher than the continuous wave laser,
resulting in high absorption.
Laser power is a critical parameter to be defined for different purposes.
Laser Processing of Magnetic Materials 71

For continuous wave power density. density, watt per unit area is important for laser
surface treatment. For pulsed mode peak, power is defined to be the energy
per pulse (joule) divided by the temporal length of the pulse. Other related
terminologies are: Average energy density (J/ cm 2 ) . average power density
(W/
(W/cmcm 2 ) ,. peak power density (W/ cm 2 )) ..
(W/cm
Monochromaticity referrs to the property of having all the photons in the
laser beam with exactly the same wavelength. Monochromaticity is very
important for laser Iithography
lithography for producing a clean interference pattern.
Monochromaticity could also become the reason for achromatic aberration of
the optical system. In reality no lasers are exactly monochromatic.
Monochromaticity is described by wavelength line-width which covers the
wavelength in the laser.
Coherence has two meanings: Temporal coherence refers to the relative
phase or the coherence of the two waves at two separate locations along the
propagation direction of the two beams. It is also referred to as longitudinal
coherence. If we assume that the two waves are exactly in phase at the first
location, then they wi willII sti
stillII be at least partially in phase at the second location
up to a distance 'e' where I'eis c is defined as the coherence length. Spatial

coherence refers to the property of having all the photons transmitting in the
some direction. Spatial coherence defines how far apart two sources can be
located in a direction transverse to the direction of observation and still exhibit
coherent properties over a range of observation points. It is also referred to as
the lateral coherence. Both temporal coherence and spatial coherence are
important for lithography.
The beam mode refers to the intensity profile of the output spot. Most
commercial lasers have a TEM oo mode which has a Gaussian profile in cross-
section. TEM is the abbreviation for transverse electromagnetic mode. Images
of several mode distributions are shown in Fig. 3. 1a as they might be observed
if a laser beam were projected onto a screen. These mode patterns occur for a
completely symmetric laser medium in the directions transverse to the
direction of travel of the laser beam. Modern lasers can integrate several
modes to generate ....... much smother beam intensity as shown in Fig. 3. 1b .

3.2.2 Lasers Available for Materials Processing


CO2 lasers can generate the highest power and operate in the wavelengths
between 9. ~m. either by continuous wave or pulsed mode. Both
9.44 and 10. 6 IJm,
modes are high-average-power and high-efficiency laser systems.
Commercially available continuous wave CO 2 lasers range in power from 6 to
10, 000 W, and custom lasers are available at even higher powers. CO 2 lasers
are widely used in diverse commercial applications such as marking of
electronic components, wafers, and chips.
chips, marking on anodized aluminum,
trophy engraving, acrylic sign making, rapid prototyping of 3 dimensional
72 Vi Liu et al.

-O$@
-06)@
TEM oo
oo TEMOI'
TEM01' TEM IO
lo TEM II *
TEMI1'

(a)
-
T EM ol
01 TEM o2
~:c
TEM o) TEM o4

OJ r---,-----.--------,-----,--,-------,------.------,
,----,----,------,-----.--,-------,------.------,

.2
.f' 14% "00" mode
5~ 0.2 f-----+--I--+---+-----f--+--+---\--+---i
f------+--I--+---+----f---f---+---\--+----j \4% "10" mode
14%
.55 ...
36% "0\"
.6. "01" mode
-0
36% "02" mode
.~ + "Q"Mode
"Q" Mode
-;;
<;;
~ 0.\
0.1 f------+------:..Ikc--~f_I_--+-\A_V-
f-----+-----:~c__~f+----f-\A_1r-___:i~-~---i
~---+-----j

-1.5 -1.0 -0.5 o 0.5 1.0 1.5 2.0


Position
(b)

Figure 3.1 (a)


Ca) Laser beam image as if they were seen. (b) Cb) Laser modes profile in a
CO2 laser. The overall intensity is the sum of four modes which is much more uniform than
a single mode.

models, cutting of ceramics, textiles, and metals, carpet, saw blade,


cutting, drilling, thin film deposition, wire stripping, surface treatment,
surface alloying, and cladding. Because of the high power, it has the potential
for magnetic materials processing such as thin film surface treatment.
Dye lasers are produced in liqUid
liquid gain media. The word dye refers to dye
molecules including polymethine, xanthene, coumarin and sclintilator.
sci intilator. The
laser gain media are made of solutions of dye molecules in water, alcohol, or
ethylene glycol. The dyes have very broad emission and gain spectra with
tunable laser output and short-pulse laser output. A typical dye laser can
operate over a wavelength variation of 30 - 40 nm. Tunable laser output over
wavelengths ranging from 320 to 1200 nm are available thbyough using various
dyes. Dye lasers can be continuous or pulsed. The output of continuous dye
laser can go up to 2 W. Pulsed dye lasers pumped by other lasers (such as
excimer, nitrogen, or frequency-multiplied
frequency-multi pi ied ND:
NO: YAG lasers) can produce
output pulses up to tens of millijoules in a 10-ns pulse at repetition rates of up
to 1 kHz.
Excimer lasers are pulsed lasers incorporating electronic transitions within
word""excimer"
short-lived molecules. The word excimer" is an abbreviation of the phrase
"excited dimer". Such lasers are most often composed of the combination of a
rare-gas atom (such as Ar, Kr, or Xe) and a halogen atom (FI, CI, Br, or I).
The excimer molecules exist only as excited molecules because the ground
Laser Processing of Magnetic Materials 73

state is extremely short-lived,


short-l ived, due to the repulsive force between the two
atoms of the molecule in its ground state. The major excimer lasers are XeF at
353 nm, XeCI at 308 nm, KrF at 248 nm, ArF at 193 nm, and F2 at 153 nm.
The pulsed output of these lasers has a duration lasting from 10 to 50 ns with a
pulse energy of O. 1 - 1 J, and can be operated at a repetition rate of up to
several hundred hertz.
X-ray lasers use hot plasma such as Ar8 +, + , Ti 12 +, Fe 16 +, Zn20 + ,
Ge22 + , 8e24 + , Y29 + , Ag37 + ... as emitting media which is produced by
focusing a high power pulsed infrared-laser on a solid target. The wavelength
varies from 60 to 1 nm corresponding to the atomic number from light element
to heavy element. The pulse averaged power is at the order of tens of
megawatts which offer a brightness of 7 - 8 orders higher than ordinary X-ray
sources. The weakness of such lasers is that the repetition rate is very low
because of the pumping laser cooling time which requires from minutes to
hours.
Beam del iver system: a laser beam can be shaped/del ivered by lenses,
mirrors and fibers. An optical system can be used for the purpose of getting
the desired spot size, uniform intensity profile or interference pattern.
Two ways have been used to smoothen the intensity profile. An integrated
mirror can be constructed to generate an averaged beam profile as shown in
Fig. 3.2. TEM o1 * beam, for example is reflected by an integrated mirror which
is made of many small segments. These segments focus the beam at a
distance c, generating a square pattern. In each square pattern the intensity
in a position (x,
(x , y) is contributed by all the segments in the integrated mirror.
Oscillators can be connected to the two mirrors to generate x-direction
oscillation and y-direction oscillation.

Flat mirror

Integration mirror

Figure 3. 2 Optics of integrated mirror.


74 Vi Liu et al.
Yi

3.2.3 Laser Materials Interaction


When a laser beam impinges on the surface of an object, the laser energy can
be reflected, transmitted and absorbed. The energy absorbed is used for
materials processing. The absorbtion can be described by
A =- dill = z(ao
z(O:o + O:ind
aind + {3J)
f3f) (3. 1)

ao is the linear
where z is the path length of the laser beam, I is the intensity, 0:0
absorption coefficient associated with single-photon absorption, O:ind aind is
associated with laser induced defect. {31f31 defines the linear increase of laser
absorption A with I and originates from the two-photon absorption mechanism.
When the total energy of two photons exceeds the band gap, the excitation of
the electron can occur by absorbing the energy of two photons. The linear
dependence of absorption on intensity has been observed in BaF 2 (Eva and
Mann, 1996).
The absorption of laser beam by materials is quite complicated in laser
materials processing. This is because the radiation by a laser beam will
introduce defect structure in the materials, causing different absorption
mechanisms. Therefore, for each case, the absorption has to be determined
experimentally.

3.3 Laser Surface Treatment of Magnetic Thin Films

Laser surface treatment is accomplished by exposing the surface of a specimen


under laser radiation. Laser surface treatment of magnetic thin films is suitable
because the film is very thin in the order of 10-
10 - 30 nm which is a small fraction
of the laser wavelength 248 - 10, 000 nm. The laser energy absorbed within
10,000
this layer has very small variation. The mechanism for properties improvement
Iies in the phase transformation accompanying the rapid heating and cooling.
lies cool ing.
One example of successful application of laser surface treatment has been the
writing of data on magneto-optical recording media. First the film is
magnetized in one direction. A proper opposite magnetic field is appl ied. As
applied.
the film has higher coercivity, the magnetization will remain stable. A laser
beam is focused on a recording bit to raise the temperature of that area above
Curie temperature. Under the applied magnetic field only this area will
undergo magnetic reversal, and a recording bit is thus written.
Laser Processing of Magnetic Materials 75

3.3. 1 Precise Grain Size Control of Magnetic Thin Films

In conventionally fabricated nanostructured materials, the grain size covers a


wide range. This is unfavorable for the case of magnetic recording media.
High density recording at the level of 100 Gb/in 2 requires the grain size to be
very small as well as uniform (Zhou et ai.,
al. , 1999). Larger grain size than the
designed grain size will increase the roughness of the magnetic transition zone
and therefore increase noise, while small grain size will cause a thermal
stability problem since the K u V/ kT == 50 is a criterion for the minimum volume
of a grain. It is then very desirable to be able to control the grain size at d =
do + c where d is distance for a molecule to transter from the parent phase to
the grain, and c is the tolerance for grain diameter deviation form the
designed value do. The following describes a process for precise gain size
control by laser surface treatment of magnetic films.
The grain size of a phase is the result of nucleation and grain growth.
Therefore, the grain size in a conventional material has a range. Several
factors can affect the final grain size. First, and most importantly, the earl
earlier
ier
formed nuclei undergo grain growth while other nuclei are being formed. This
process continues until the phase transformation finishes. Second, the size of
the critical nucleus is a function of temperature and nucleation site. The
temperature could vary with specimen position and with time. The nucleation
site could be within the grain, at grain boundaries, at interphase interface
boundaries, at dislocations, etc. For simplicity,
simpl icity, here we only deal with
homogeneous nucleation. Mathematically considering the phase transformation
from the parent phase a ex to the new phase f3, the nucleation rate N, defined to
be the nuclei formed per unit time and unit volume, is

..
= No exp
N =
(t.Gt)
(boG t ) (t.G*)
(boG)
- K B T exp - K B T (3.2)

No = (Nmv) (3.3)

where N m is the total number of molecules in unit volume, v is the frequency of


effective vibration leading to an atom transfer from the parent phase to the
t.G,
embryo, boG t is activation energy for a molecule to transfer from the parent
K boG
Soltzman' s constant, t. G * is free energy change
phase to the embryo, K BB is Soltzman's
for the formation of a critical embryo and is expressed, for the case of
spherical embryo, as

(3.4)

where t.G
boG is the free energy change per .unit unit volume during phase
transformation, and y is the interphase energy betw.een
between the embryo and the
embryo has the radius rr*
parent phase. The corresponding critical embryohas
76 Vi Liu et al.

, 2y
(( == 6.G
~G' (3.5)

The growth rate of a grain towards the parent phase G is

G == Go ex p [ - 6.K~t
~K~tJ[] [11 - exp( - :B~ )]
:B~) ] (3.6)

Go = a'
a * vd (3.7)

where a' is a factor due to lattice structure, d is the distance for a molecule
to transfer from the parent phase to the grain.
Next we deduce the transformed volume fraction as a function of time
under isothermal condition. Let the transformation occur in the time interval
0< t< T in a unit volume. The nuclei formed at t == t will grow until t == T, the
f3 formed at time dt throughout the ex phase is given by
volume fraction of {3
= (4lT/3)(GCT
dV s = (41T/3)( G CT - t) + ('
(* Ndt.
)3Ndt.
)3 (3.8)
Integration of Eq. (3.8) from t = 0 to t = T yields
V =
Vs
= 1TlTN[(
N [ ( GT + r' )4
(*)4 - r'
(*4J
4]
(3.9)
3G .

At time t = T, let the volume fraction of the parent phase that has
transformed to the ~ phase be V V ~, and the untransformed fraction is 1 - V ~ .
Then the calculated V ss by Eq. (3.9) has two parts: one is in V ~ which is
already transformed and the other in 1 - V ~ which is untransformed. V ss is
therefore called superficial transformed volume fraction. The transformed
volume fraction at t = T is thus given by
dV~ =
= dV s(1
(l s - V~). (3.10)
Integration of Eq. (3.10) yields
Eq.(3.10) V~=l-exp(-Vs)
V~= 1-exp(- Vs ) or

- 1_ f _ IT
{_ GT + ('
1T N [ ((GT (* )) 4 - (
(* J}
* 4] }
(3. 11)
V ~~ - ex p \
- exp 3G .

The duration of phase transformation T from Eq. (3. 11) is

T
[ - 3GI~~ - V~) + (*4
[-
= -=----'---------=----
= -==--------------=----
(* r- 1

(3. 12)
G
Next we consider the isothermal condition. The minimum grain size (* is
= T and the maximum grain
formed at the end of phase transformation at time t =
= O. The grain growth 6.(
size (m is formed at time t = ~( is

(3. 13)
(3.13)

In order to reduce the grain size difference, both G and T should be reduced.
G is a complicated factor affected by both material composition and
temperature. The most straight forward approach is to reduce T, which can be
done most effectively by increasing N. By inspecting Eqs. (3.2) and (3.4) we
Laser Processing of Magnetic Materials 77

notice NtV increases with small interphase energy y and large energy change
/::"G. The volume free energy change /::,.G could be approximated by
/::"G,
/::,.G = /::"s(T
/::"s( T - T
To)
e) (3.14)
(3,14)

where /::"s is the entropy change accompanying the phase transformation, and
T e is the equilibrium temperature Of
T. of a phase and 13. The laser radiation has the
and~.
ability to bring the thin film to a temperature suitable for high NtV and small G
and T.T, Our numerical calculation has shown that the phase transformation can

(a)

(b)

(c)

Figure 3.3 TEM images of Co-Sm


Co-8m films. (a) conventional annealing (b) radiated by
CO2 laser for 3 seconds and (c) HRTEM image of (b).
78 Vi Liu et al.

be finished by laser processing in 1 to 3 s by increasing the temperature to


1000 'c while 30 min are needed for conventional heat treatment at 600 'c .
Figure 3.3 compares the transmission electron microscopy (TEM) images of
CoSm films processed by (a) conventional annealing (b) and (c) radiation
from a CO 2 laser for 3 s. Laser treated film has very small and uniform grain
size because the heat input can be precisely controlled. Therefore laser
Ims, although not used in production of
surface treatment of magnetic thin fifilms,
recording media at'
at present, represents a unique potential for nanostructure
design.

3.3.2 Alignment of Magnetic Grains by Laser Annealing


High signal level in a recording media is expected if a uniaxial texture with
c-axis al igned in one direction can be achieved. In the past, control of
substrate temperature during film deposition has been used to simulate (001)
texture and (110) texture of Cr (Feng et al., aI., 1994; Laughl in and Wong,
1991). The magnetic layer of the Co alloy will epitaxially grow on the Cr film
(Mirzamaani et al. , 1990, 1991; Min and Zhu, 1994). A single crystal of Cr
with (112) plane surface can be used to grow uniaxial anisotropy Co alloys
aI., 1995). Annealing of an as-deposited Fe/Pt multilayer will form
(Liu et ai.,
L 10 structure, FePt. Depending on the annealing temperature, two kinds of
textures, the c-axis perpendicular to film plane or c-axis in-plane texture, can
be achieved (Liu et ai.,
aI., 1998, 1999). In the case of c-axis in-plane texture
the c-axis is randomly distributed in the film plane.
Alignment of magnetic grains by laser-induced phase transformation under
magnetic field. Laser-induced phase transformation provides the unique
opportunity for achieving a desirable texture because laser energy has the
advantage to be easily delivered to the film placed in a magnetic field.
Figure 3.4 shows our experimental set-up for laser surface treatment of
magnetic films. Ar gas is used to protect the film from oxidation. The as-
deposited multilayer Fe/Pt film has an face centered cubic (fcc) structure
which is magnetically soft. Heat treatment at 600'C will introduce the phase
transformation from fcc structure to L 10 structure which is a hard magnetic
material. In our experiment we found evidence that uniaxial texture can be
achieved by annealing an as-deposited Fe/Pt multilayer under an applied
magnetic field. In the process of phase transformation from fcc Fe/Pt
multilayer to L 10 structure, FePt the c-axis has a tendency to be aligned
parallel to the magnetic field. Figure 3.5 shows the hysteresis loops.
Magnetization reversal occurs in a much narrower region in Fig. 3. 5a than in
Fig. 3. 5b. This is an indication that there are more grains with magnetic easy
axis in the direction parallel to the magnetic field. The magnetic field is
generated by a permanent magnet with a field of about 1.2 T. Alignment effect
can be enhanced by applying stronger magnetic field.
Laser Processing of Magnetic Materials 79

Laser

PYmm"e~
PYromete~

ArGas c=]
C]
Magnet

Figure 3.4
3. 4 Experimental set-up for laser surface treatment of magnetic films.

rr
.: I .:f
-1---L----i--
: I :
: I :

-I L I..J
-5 o 5
H(kOe)
(a)

-J--L-l--
!~
--+--L---l--
.r
T
. I j
I !j
-I
-)
./;
.</ V
~

-5 0 5
H(kOe)
(b)
Figure 3.5 Hysteresis loops of a Fe/PI' film (a)
Ca) measured parallel to the magnetic
filed and (b)
Cb) perpendicular to magnetic field.

3.
3.44 Interferometric Laser Lithography

Periodic magnetic nanoarrays are thought to be useful in


in patterned magnetic
recording media and magnetic random access memories (MRAM). Several
80 Vi Liu et al.

groups have used interferometric laser lithography (ILl) to make large area
arrays of magnetic dots of about 200 nm dot-spacing (Schattenburg et al. ,
1995; Spall al. , 1996). The minimum spacing achievable by ILL is ~A /2,
as et a!.
Spallas
where ~A is the wavelength of the employed laser. Recently dot-spacing down
to 100 nm nanoparticle arrays has been fabricated by using an achromatic
interferometric lithography system (Savas et al., 1996). Until now, the
application of ILL in patterning magnetic nanostructures has been divided into
two categories: one is multi-step process which involves the resist, etching
patterning/anneal ing. In the latter
and subsequent lift-off; the second is direct patterning/annealing.
case the film is exposed to interferometric laser directly without the assist of
resist. We will briefly summarize these two processes and the corresponding
magnetic properties of the patterned magnetic arrays in the following section.

3.4. 1 Multi-Step ILL


Briefly, a resist such as polylmethyl-methacrylate (PMMA) coated on a Si
wafer is exposed to an interference pattern. Then it is exposed a second time
after being rotated 90 with respect to the first pattern. After developing the
exposed layer, two-dimensional arrays of resist holes are generated.
Magnetic materials are then evaporated into the holes. The subsequent lift-off
process removes the rest of the resist and leaves the magnetic dot array.
Some groups have produced Co and Ni nanodot arrays by this method.
Fernandez et al. (1996) studied the magnetic properties of Co dots
patterned using magnetic force microscope (MFM). The authors found that for
Co dots with diameter of 100 nm and thickness of 40 nm, the dot is uniformly

(a) (b)

Figure 3.6 MFM image of an array of Co dots: (a) Co dots with diameter of 100 nm
and height of 40 nm. The dipole signature of these dots indicates the moments are in-
Cb) Co dots with diameter of 70 nm and height of 100 nm.
plane; (b) nm, The symmetry of the
images shows that the moments are out-of-plane (Fernandez 1996).
CFernandez et al. 1996),
Laser Processing of Magnetic Materials 81

magnetized in a direction predominately parallel to the substrate, as


characterized by the dipole-like MFM image. While for Co dots with diameter
of 70 nm and thickness of 100 nm, MFM images show that the dots are
light in the middle. The authors
circularly symmetric and are either dark or Iight
suggested that these types of Co dots are single-domain and magnetized in a
direction predominantly normal to the plane of the substrate. Figures 3. 6a, b
show the MFM images for the two types of Co dots described above.
al. (1999) studied the magnetic properties of Ni "pillars" or
Savas et a!.
"pyramid" by being electrodeposited or evaporated into ILL generated
templates. The Ni pillars have heights of 115 nm and widths of 58 nm,
De. The saturation
exhibiting out-of-plane easy axis with coercivity of 700 Oe.
magnetization is 0.5 memu/cm 2 . For Ni pyramids, which have 35 nm height,

. ...,
~~,.~.&:'tJ'"
4 '.
_ ---'1-1:I ~ ~.:EIo'~'"
33ilP'lJ"'-~'IQ
ID
qP ...::-"'0
CCI I:CI
~
[lJOl
tn ....
.
. . ,' Q
~E
ccfJ
C cP 2 ,, '.'
2._ D

...
I:POD
tP
~ o , O

E
Q)
rj'rfJ "
1:IrJ'r5'-
tJ ~.:",sJJ
I
&Il
"r.
b--2000
2S
000
. -1000
ca
%
q,
1:10
DD ,
~"1"
...
~
~
ctf
c$P
r9
~
1000 2000
20 00

i
o
, <I
<J
.~
CO
CD
.....
.'
...: . :
,t.L. ~cD!b
.' ,.,g1P-
~2
Cb

~ 8 ..D~'~-3
j

.. ..
~.....4.~'1l'
~","4.~~..
'"
-4
-4 Field in plane
co Field out of plane
Applted
Appiled field (Oe)
Measured at 10K
(a)

44
.....".,.
1lo.
r;;J.
D .0 . . . . . . .D. .'
D
~~.~:....r..: .......:.
D

d' ~"~~DUo'll1l
.. "'~7J'"
33
.-. ~"
2~"
~
g,
E
1
~
~E
2~..
It 1
:-.4
:-.~
I ~,'
!
a
D'
c

.
Q)

b- -2000
2000 -1000 1000 2000
"I'
o"l.j
2S 0- 1

"::Ec
Q)

E
o
1:1 i!...
:i!...
'11!:2
"'-2
o
~ ..~!~~~~3
..
..
~~!.~~~3 ' ' gc
.
.,'.
:. -4 Field in plane
o Field out of plane
cField
Applied field (Oe)
Measured at 300 K
(b)

Figure 3.7 SQUID hysteresis loops for an arrays of evaporated Ni pyramids at 10 K (a)
and 300 K (b) (Savas, et al. ).
82 Vi Liu et al.

30 nm diameter and 70 nm inter-particle spacing, there is an out-of-plane easy


axis at 10K. Figure 3.7 shows the hysteresis loops for the Ni pyramids
measured by a superconducting quantum interference divice (SQUID)
magnetometer.

3.4.2 Direct Patterning Using ILL


Direct patterning using ILL is a non-contact, straight-forward method. Its
simplicity cleanl iness is highly desirable for future magnetic
simpl icity and relative cleanliness
nanostructures and devices. For this patterning method, short-pulsed (down to
10 ns) interfering beams ofaXeCI laser (308 nm) are used to expose Co-C
films (Zheng et al., 2001). The laser beam was split into two beams of
approximately equal intensity and then recombined, generating an interference
pattern due to a periodic modulation of the Iight
light intensity. The interference of
the four beams leads to the formation of two-dimensional arrays of modified
regions having a minimum lattice periodicity equal to half of the radiation
wavelength. The size and shape of the regions depend on the incident angles
of the beam, pulse duration, and flux. Co-C films are chosen in this
exploratory study because they form a simple system containing immiscible
magnetic (Co) and non-magnetic (C) phases. The metastable Co carbides
decompose easily into Co and C upon annealing (Hayashi et al., 1996; Yu
et al.,
al. , 1999; Konno and Sinclair, 1994; Delaunay et al., ai., 1997). Magnetic
properties have been studied in laterally homogenous films in both the as-
deposited and annealed states (Hayashi et al. , 1996; Yu et al. , 1999). One
can therefore expect that the behavior of laser-quenched alloys is similar to
that of thermally treated Co-C films. The Co dot arrays are formed by phase
separation of Co or Co-rich clusters under the influence of laser intensity
maxima which provide local annealing. Figure 3.8 shows the AFM image of a
patterned Co-C fifilm. Per iodic dot arrays are generated in a large-scale area
1m. Periodic
20.0nm
20.0 nm

10.Onm
IO.Onm

Onm

Figure 3.8 AFM image of 250 nm x 500 nm periodic arrays of patterned Co-C films. The
scan size is 30 IJm.
Laser Processing of Magnetic Materials 83

with a dot diameter of 250 nm and center-to-center spacings of 375 nm x


750 nm. The observed topographical contrast from AFM picture is thought to
reflect the surface lifting in interference maxima by phase changes in carbon.
It was found that the magnetic properties of Co nanodot arrays depend on
the power of laser being employed. The formed dots can be embedded either
in paramagnetic matrix or in a weakly magnetic matrix. The as-sputtered Co-C
film is paramagnetic at room temperature. Figure 3.9 shows the temperature
dependence of the magnetization of the as-sputtered Co-C film at 100 Oe. The
hysteresis loops at 5 and 300 K are shown in the inset. The as-sputtered film
exhibits a ferromagnetic hysteresis loop at low temperatures, but the collapse
of the magnetization far below the Curie temperature of Co (1388 K) indicates
that this hysteresis is due to very small superparamagnetic Co particles. After
laser patterning, magnetic hysteresis is observed at 5 and 300 K (Zheng
et al. , 2001), and in both cases the coercivity is of the order of 100 Oe.
2.5
3
--50K
S- 2 ----300K
S' ----- 300 K
aE
2.0
.

Q)

'0
"b
x
'-'

s-
S' \.
\
a
E
~Q) 1.5 l.
o
x
~
'-'
\
c:
.s
.9

.,
.~
Q)
c:
1.0 - 3 'L...J._--'-----..J'--------'-:-_~
-3 --'_---L_--l-_L--'-_
-1000 -500 0 500 1000
fn
on
co
Field (Oe)
2'"
::E
0.5

o .................
o 50 100 150 200 250 300
Temperature (K)

Figure 3.9 Temperature dependence of the magnetization of the as-sputtered Co-C film
measured at 100 Oe by SQUID. The hysteresis loops at 5 and 300 K are shown as inset.

Figure 3. 10 shows AFM (Fig. 3. lOa, b) and the corresponding MFM


(Fig. 3. 10c, d) images of Co-C composite films after they were exposed to a
moderate-intensity interferencing laser beams. In Fig. 3.10c, d, a
appl ied in situ in directions parallel and
perpendicular field of 1700 Oe was applied
anti-parallel to the film-plane normal, respectively. The bright and dark
contrast in MFM images corresponds to the strength of the stray-field gradient
84 Vi Liu et al.

on the sample surface. The Iight color indicates the force between the sample,
and the MFM tip is repulsive. The clear MFM contrast between the two

..
directions of the applied field indicates that the formed dots are magnetic,

....,--.,-;-;-;."
.....
while the area between the dots is non-magnetic (Zheng et al. , 2001).

...... .........
'.,-;-;-;
,-

-

.. .
i......
-..........
......, _.1
."..-.'
,


1-
..
Ii.e,.-.-.'
~' '.,11

..........
.....!.
- .. A.c- ......
, -
I

-,'
, I1
. . . . . . . . . . e'l

.......... "'"'1I~
. . . . . . . . . -,

-.-.-.-.-
.........
'-.-;.;-;-;. .
(a) (b)


...... ,


..... .....

...... -
.
..
.. .
II . _

..... .....
-
, .
#





~
a..;.
.-r=
~
(c) (d)

Figure 3.10
Figure3.10 AFM (a) and (b) and MFM (c) and (d) images of a periodic array of
patterned Co-C films after exposing the film to interference laser. A perpendicular field of
1700 Oe was applied parallel (c) and anti-parallel (d) to the film-plane normal during
scan.
scan,

However, if the film is exposed to high intensity of laser, the area


between dots can be transformed into magnetic state (Zheng et al. , 2001).
Figure 3. 11 shows a MFM image of the patterned Co-C film under higher laser
0
power. A 180 180' domain wall was observed, as evidenced by the reversed
black-white contrast of the dipole-like shape of the dots. The domain pattern
indicates the presence of interdot exchange interaction via the ferromagnetic
matrix,
matrix. The reason for the induced ferromagnetic character of the matrix is
most likely due to the comparatively high laser intensity (Zheng et al. , 2001).
Suppose there is a threshold value of laser power, E t . Only when the laser
intensity is higher than E t can the film be made magnetic due to the Co atoms
re-arrangement or phase separation upon anneal ing. At moderate laser power,
annealing.
e.g., Fig. 3.10, only dot regions formed under the laser intensity maxima
whose power is higher than the threshold become magnetic while matrix under
Laser Processing of Magnetic Materials 85

the laser intensity minima remains paramagnetic. If the employed laser power
is higher than EE"t , both dots and inter-dot regions become ferromagnetic,
which is seen in Fig. 3.11.

'11111~~I~~~~~
~ 180
0

Domain
180
wall

Figure 3. 11 MFM image of the patterned Co-C film under higher laser power.

3.5 Concluding Remarks

In this chapter we have given a review of the principle, important aspects,


and several examples of laser processing of magnetic materials. Methods such
as laser lithography have been used for a long time and offer unique
capabilities. Laser surface treatment of magnetic thin films offers the
advantages of high efficiency, clean, precise location and time-duration
control. When combined with applied
appl ied magnetic field, grain size control and
texture control are possible which might be suitable for processing of magnetic
recording media and are not easily realized by any other methods. To
fabricate useful nano-dots at length scale of 5 to 10 nm, high peak power,
short pulse and short wavelength lasers tunable around 10 nm must be
developed.

References

Delaunay, J. -J.,
-J. , T. Hayashi, M. Tomita and S. Hirono. J. Appl. Phys. 82:
2200 (1997)
Eva, E. and K. Mann. In: M. Morin, A. Giesen ed. Third International
Workshop on Laser Beam and Optics Characterization. SPIE, 2870: 476
86 Vi Liu et al.

(1996)
Feng, Y. C.,
C. , D. E. Laughlin and D. N. Lambeth. J. Appl. Phys. 76: 7331
((1994)
1994)
A.,, P. J. Bedrossian, S. L. Baker, S. P. Vernon and D. R.
Fernandez, A.
Kania. IEEE Trans. Magn. 32: 4472 (1996)
Hayashi, T., S. Hirono, M. Tomita and S. Umemura. Nature 381: 772
(1996)
Konno, T. J. and R. Sinclair. Acta Metall. Mater. 42: 1231 (1994)
Konno,T.
Laughlin, D. E. and B. Y. Wong. IEEE Trans. Magn. MAG-27: 4713 (1991)
Liu, Y. , D. J. Sellmyer, B. W. Robertson, Z. S. Shan and S. H. Liou.
IEEE Trans. Magn. MAG-31: 2740 (1995a)
J. P. , C.P.
Liu, J.P., C. P. Luo, Y. Liu, D. J. Sellmyer. Appl. Phys. Lett. 72: 483-
D.J.
485 (1998)
Liu, P.,
P. , Y. Liu, R. Skomski and D. J. Sellmyer. IEEE Trans. Mag. 3241-
3246 (1999)
Liu, Y. , J. Koch, J. Mazumder and K. Shibata. Metall. Mater. Trans. B 25 :
425 - 434 (1994a)
Liu, Y. , J. Mazumder and K. Shibata. Metall. Mater. Trans. B 25: 749-
759 (1994b)
Liu, Y. , J. Mazumder and K. Shibata. Acta Metall. 42: 1763 - 1768 (1994c)
Liu, Y. J. , Mazumder and K. Shibata. Acta Metall. 42, 1755 - 1762 (1994d)
Liu, Y. , J. Mazumder and K. Shibata. Metall. Mater. Trans. A 26A, 1519-
.1533
1533 (1995b)
Mazumder, J., J. Koch, K. Nagarathnam, J. Choi. Fabrication of 3-D
Shapes by Laser Aided Direct Deposition of Metals. 1997 Fall Meeting of
the Materials Research Society, Boston, MA, December 1- 5 (1997)
Min, T. and J. G. Zhu. J. Appl. Phys. 75: 6129 (1994)
Mirzamaani, M., K. Johnson, D. Edmonson, P. Evett and M. Russak. J.
Appl. Phys. 67: 4695 (1990)
Mirzamaani, M., C. V. Jahnes and M. A. Russak. J. Appl. Phys. 5169
( 1991)
Savas, T. A., M. Farhoud, H. I. Smith, M. Hwang and C. A. Ross. J.
Appl. Phys. 85: 6160 (1999)
Schattenburg, M. L., R. J. Aucoin and R. C. Fleming. J. Vac. Sci.
Technol. B 13: 3007 (1995)
Spallas, J. P., R. D. Boyd, J. A. Britten, A. Fernandez, A. M. Hawryluk
J. M. Perry and D. R. Kania. J. Vac. Sci. Technol. B 14: 2005 (1996)
Yu, M., Y. Liu and D. J. Sellmyer. J. Appl. Phys. 85: 4319 (1999)
Zheng, M. , M. Yu, Y. Liu, R. Skomski, S. H. Liou, D. J. Sellmyer, V. N.
Laser Processing of Magnetic Materials 87

Petryakov, Yu. K. Verevkin, N. I. Polushkin and N. N. Salashchenko.


Appl. Phys. Lett. 79: 2606 (2001
(2001a)
a)
Zheng, M.,
M. , M. Yu, Y. Liu, R. Skomski, S. H. Liou, D. J. Sellmyer, V. N.
Petryakov, Yu. K. Verevkin, N. I. Polushkin and N. N. Salashchenko.
IEEE Trans. Magn. 37 :2070 (2001b)
Zhou, H.,
H. , H. N. Bertram, M. F. Doerner and M. Mirzamaani. IEEE Trans.
Mag-35: 2712 (2001)

This research was supported by NSF, DOE, ARO, AFOSR, NRI,


NR I, CMRA, and NSF-MRSEC
CDMR-0213808) .
4 Processing and Properties of Nanocomposite
Nd2Fe14B-Based Permanent Magnets
NdzFe14B-Based

Satoshi Hirosawa

4. 1 Introduction

The nanocomposite Nd2Fel4 Fe14 B-based permanent magnets was produced by


researchers at Philips Research Laboratories for the first time ( Buschow
et al.,
ai., 1988; Coehoorn et ai., a!., 1988). The materials comprised of Fe3 B,
Nd2Fel4 B, and a small amount of a-Fe.
ex-Fe. The content of Nd in these magnets is
about one third of conventional Nd2Fel4 Fe14 B-based magnets, and hence a lower
material cost was anticipated. Later, a semi-reversible demagnetization
behavior originating from nearly reversible rotation of magnetic moments in the
magnetically softer components was noted, and the term "exchange-spring
magnet" was coined by Kneller and Hawig (Kneller and Hawig, 1991). More
recently, Skomski and Coey, pointed out the possibility of realization of
magnetic energy products exceeding 1 MJ/m3 (Skomski and Coey, 1993),
2. 5 times greater than the value (444 kJ/ m3) achieved in the
which is about 2.5
best Nd2Fe
Fe14 B-based magnet reported up to this date (Kaneko, 2000). These
14B-based

experimental and theoretical findings have kindled intense research on


nanocomposite permanent magnets.
The magnetically hard component of nanocomposite permanent magnets
can be selected from known hard magnetic phases such as FePt, SmCos ,'
Sm2 Fell
Fel7 N3 , and Nd2Fel4
Fe14 B. However, Nd2Fe14 Fe 14 B is practically the only phase
ex-Fe or Fe3 B phase of a large
which can coexist in the fully developed crystalline a-Fe
magnetization. Thus, a-Fe/N~
ex-Fe/N~ Fe14
Fel4 Band Fe3 B/N~
B/Nd2Fel4 B nanocomposites have
been most intensively studied. This chapter describes processing and
properties of Nd-Fe-B nanocomposite permanent magnets.

4.2 Basic Considerations on Processing Techniques

According to theoretical investigations (Kneller and Hawig, 1991; Skomski


and Coey, 1993; Schrefl et a!.
al. , 1994; Fukunaga et al.
a!. , 1996), the grain size
Processing and Properties of Nanocomposite Nd, Fe" B-Based . . . 89

of the magnetically softer component in an Nd2 Fe14 B-based nanocomposite


2 Fe,4

magnet must be smaller than about 20 nm so that the magnetic moments of this
component can have strong intergranular exchange coupling with the hard
magnetic component (Nd 2 Fe,4 B). In turn, their direction is effectively locked
(Nd2Fe14B).
to prevent unfavorable rotation of magnetization therewithin. Such coupling
coupl ing is
realized only when there exist atomic contacts among the component phases.
Since the coupling energy scales with the area of interfaces formed between
the two magnetic components, a nanocomposite permanent magnet must have
a structure such as schematically illustrated in Fig. 4. 1.

HighK 1
HighK]

D)
[1) High M s

IOnm
lOnm

4. 1
Figure 4.1 Illustration of "ideal" structure of a nanocomposite permanent magnet.

A nanocrystalline material may be produced from a precursor material


which has free energy larger than the desired nanocomposite phases at a
processing temperature. One of most the practical methods to create a
structure like the one shown in Fig. 4. 1 is to make use of phase separation of a
homogeneous solid solution at a temperature where atomic diffusion is still
slow. In the history of permanent magnetism, the Fe-AI-Ni-CoFe-Al-Ni-Co magnets
(Alnico) and the 2-17-type Sm-Co magnets (Senno and Tawara, 1974) are
examples produced by phase separation reactions. Another practical process
to create a nanocomposite magnet is crystallization of an amorphous alloy
produced by rapid solidification of a melt or by sputter deposition. In this
case, phase separation is accompanied with transformation of the amorphous
to crystalline phases. During phase separation or crystallization reactions,
constituent elements have to diffuse to form concentration fluctuations which
eventually build up to form the magnetically hard and soft components. The
three-dimensional periodic arrangement of these components as illustrated in
Fig. 4.1
4. 1 develops more or less automatically in these reactions.
The reaction starts when temperature is high enough to allow atomic
diffusion of constituent elements. On the other hand, in order to avoid
unfavorable crystal growth during the reaction, temperature must not be too
high. Practical operation may be performed at temperatures ranging from
about 550 'c 'c. Figure 4. 2 shows a transmission electron
"C to about 750 "C.
picture of a Fe3 B/Nd2 Fe14 B nanocomposite permanent magnet obtained by
2 Fe,4
90 Satoshi Hirosawa

crystallizing an amorphous NdI5FenBI85CUQ2Nbl


Ndu Fen B18 sCuo 2Nb1 alloy prepared by melt-spinning
(Ping et al. , 1999b).

(c) (d)

Figure 4. 2 A transmission electron microscopic picture of a Fe3 B/Nd2 Fe"


Fel4 B
nanocomposite permanent magnet obtained by crystallizing an amorphous
Ndu Fen B'8.5 CUQ2 Nb, alloy prepared by melt-spinning (Ping et al. , 1999b).

The precursor material may also be a mixture of different phases, either


crystalline or amorphous. During an appropriately designed heat treatment,
the formation of finely dispersed structure of stable phases takes place in the
precursor material. For example, the desired amount of elemental B, Fe, and
Nd powders are mixed and mechanically milled in an inert gas atmosphere for
a few hours to produce a fine mixture of a-Fe
ex-Fe and an amorphous-like phase.
This procedure is called mechanical alloying (Benjamin, 1976). The
mechanically alloyed material is then heat treated to be transformed into an
a-Fe/NdlFeI4B
ex-Fe/Nd2Fe14B nanocomposite permanent magnet (Wecker et al. , 1995).
In some cases, a nanocomposite Nd-Fe-B may directly form from a melt
by carefully selecting solidification conditions. In a certain concentration
range, crystallization of the hard magnetic phase may be kinetically favored
over crystallization of other phases even when the latter are
thermodynamically favorable. The kinetic selection of phases takes place, for
ilistance,
instance, in an alloy of a composition in the vicinity of Ndl2Fel4
Fe14 B during rapid
sol idification from a melt CUmeda et aI.,
solidification al. , 1996). When the solsolidification
idification front
moves fast enough (at a velocity larger than about 10 mm/s) , solidification of
Processing and Properties of Nanocomposite Nd, Fe" B-Based . . . 91

Y-Fe, which requires diffusion of both Band Nd from the solidification front
y-Fe,
region, becomes kinetically unfavorable in comparison to direct formation of
Nd2 Fe14 B from the melt because of the similarity of compositions between this
phase and the liquid.

4.3 Preparation of Amorphous or Nanostructured Precursors

( 1) Technique
(1)
The chill-block rapid solidification technique as a method to produce
metallic, metastable solid solutions was introduced by Duwez et al. in 1960
(Dewez et al., 1960). Later, the twin-roller and single-roller methods were
developed to produce rapidly solidified metallic ribbons (Chen and Miller,
1970; Liebermann and Graham, 1976). The single-roller method is currently
most widely used because of its simplicity and reproducibility. In many cases
this method is referred as "melt-spinning". In melt-spinning, a stable stream of
molten alloy is ejected trough a fine orifice, the diameter of which is usually
about 0.3 - 3 mm, onto a finely-finished outer surface of a spinning roll. The
roll is usually made of Cu alloys to obtain maximum heat conductivity. By the
momentum and surface tension of the melt, it forms a small paddle on the roll
surface from which a thin ribbon or flake of the melt is drowned out as the roll
rotates. The heat of the thin melt is absorbed into the roll rapidly, and a
cooling rate which is large enough to prevent nucleation and growth of stable
phases from crystallizing is realized. Figure 4.3 illustrates basic configuration
of the single-roller melt-spinning apparatus. An industrial operation of this
method will require continuous cooling of the roll by running water inside so
that temperature of the roll surface can be maintained within a suitable range.
Ejection pressure

Molten alloy

Puddle of molten alloy

Melt-spun ribbon

Figure 4.3 Basic configuration of a single-roller melt-spinning apparatus.


92 Satoshi Hirosawa

The cooling rate depends essentially on the thickness of the melt to be


cooled, which is normally controlled by changing the revolution rate of the
quenching roller and the pour rate of the molten alloy, which is normally
controlled with the opening diameter of the orifice and the ejection pressure.
The heat transfer from the melt to the roll surface depends on various factors
such as the intrerface energy between the melt and the roll surface, size and
frequency of gas pore formation, heat conductivity of the roll material, the
heat transfer between the roll and cool cooling
ing water, and so on. After the rapidly
solidified material leaves the quenching roll, it is further cooled by radiation
and heat transfer to the surrounding gas which is normally Ar in the case where
rare earth elements are involved. Therefore, while the wheel velocity is the
most effective parameter to control quench rate, the gas pressure under which
the rapid solidification
sol idification is undertaken is also an important parameter.
(2) Rapid Solidification Process
Preparation of a nanocomposite permanent magnet by means of
crystallization of a rapidly solidified amorphous alloy involves crystallization of
multiple phases, which can be described by continuous cool ing transformation
(CCT) curves (Inoue and Masumoto, 1993). Figure 4.4 4. 4 schematically
illustrates a CCT behavior for an amorphous-forming alloy with a two-stage
crystallization process on the time-temperature plain. For curve (a), the
cooling rate is large enough to prevent any phase from crystallization and
growth, and the rapidly solidified
sol idified structure is amorphous. For curve (b), the
rapidly solidified structure is a mixture of an amorphous phase and fine
crystalline phase A, because crystallization starts after the amorphous phase
is formed and because diffusion of constituent elements to form A is already too
slow in the amorphous phase. Curve (c) is CCT for preparation of a composite
structure via a single quenching operation. To obtain a nanocomposite, the
cooling rate must be carefully optimized in order to avoid unfavorable grain
growth. This point will be more fully discussed in a later section.

Liquid
Super-cooled liquid

Time
Figure 4. 4 Schematic illustration of continuous cooling transformation (eeT)
(CCT) behaviors
for an amorphous alloy with two-stage crystallization process.
...
Processing and Properties of Nanocomposite Nd, Fe" B-Based .. 93

Although an experimental construction of a CCT diagram is by any means


difficult, it is useful to interpret the actual cooling path. For a Fe/R 2 Fel4
Fe14 B
nanocomposite, Phase A in Fig. 4. 4 corresponds to y-Fe, which transforms
o:-Fe at 1192 K, and Phase B to R22Fe14 B. Figure 4. 5 shows X-ray diffraction
into ex-Fe
patterns of Nd6 Fe88 B6 alloys melt-spun by the single-roll method under different
pressure Ar gas atmosphere with a fixed wheel velocity of 30 ml m/s.
s. In this
example, a higher chamber pressure of - 5 mm(Hg) caused a higher cooling
rate during the ribbon flow in the chamber after it departed the cool ing wheel,
and an amorphous structure resulted (corresponding to curve a in Fig. 4.4) 4. 4) .
For lower chamber pressures, a mixture of o:-Fe ex-Fe and NdFe14 B resulted
(corresponding to curve( c) in Fig. 4.4) .

2000

1500

0
.~
'in
1000
'"
I:
2~
.5'"

500

OL---'-- -L_-"'-_-----'--_ _----"-------'--J


OL----'----------'--------"'-----'--------'-------'-------'--
20 40 60 80
28(' )

4. 5 Powder X-ray diffraction patterns of melt-spun Nd6 Fe


Figure 4.5 ee 8 6 flakes under Ar
Fess
atmospheres of different pressure. The pressure is with respect to the atmospheric
ambient pressure.

In the case of Fe3 B/Nd2Fel4


Fe14 B-type nanocomposites, the situation is more
complicated than the simple case shown in Fig. 4.4. A model CCT diagram for
the Fe3 B/Nd2Fe 14B
Fe14 B nanocomposites is schematically shown in Fig. 4.6.
Figure 4. 7 shows X-ray diffraction powder patterns for Nd4Fens
Fen.5 B I85
185
alloys obtained via a melt-spinning technique with different wheel velocities
(V s ) (Kanekiyo and Hirosawa, 1998). In this case, the surface velocity of
( V s)
=20
Vs5 = ml s is fast enough to obtain the amorphous sol id. A sl ightly slower
= 10 m/s results in crystallization of Fe3B. Nd2Fe14B is present
quenching at Vs =
= 7 and 5 m/s. Therefore, there exists another C-
in alloys solidified at V s =
curve for crystallization of Nd2Fel4
Fe14 B hidden in the region B in Fig. 4.4. At a
slower cooling rate at Vs = = 3 mis, thermodynamically equilibrium phase o:-Fe
ex-Fe
(originally y-Fe, corresponding to phase A in Fig.4.4) is present. Figure 4.8
94 Satoshi Hirosawa

Time

Figure 4. 6 A model CCT Fe3 8/Nd


eeT diagram for the Fe] Fe" 8
B/Nd2 Fel' B nanocomposites (only
schematic) and continuous cooling curves corresponding to different quenching rates.

shows transmission electron microscopic images of the as-quenched materials


obtained at different wheel speeds of 3, 5 and 7 m/s (Hirosawa et al. ,
1998a). The dendritic growth of y-Fe in the specimen quenched at surface
velocity (of quenching roll) V s =3 m/s is recognized. In contrast, small
crystallites are embedded in amorphous matrix in the specimen quenched at

eu-Ka
Cu-Ka

Fe3B P=I.3kPa
P=I.3 kPa
l4 B
'".. Nd2 Fe 14
- a-Fe
a-Fe
V =3m/s _
s
.~ ~(If"III#''''''
c
2
.s ...................."'" ~'f "'..J""
Vs=7m/s

Vs=lOm/s

30 35 40 45 50 55 60 65
2en
2ee)

Figure 4.7 X-ray diffraction patterns of Nd, Fe" 5818 5 melt-spun alloys obtained under an
Nd4Fe775B185
Ar pressure of 1.3 kPa on a Cu-roll
eu-roll surface moving at different velocities (Kanekiyo and
Hirosawa, 1998).
Processing and Properties of Nanocomposite Nd, Fe" B-Based . . . 95

v s == 7 = 5 m/s, and good magnetic properties are obtained in the


m/s. At Vs =
as-cast state.

(c)

Figure4.8
Figure 4.8 TEMof
TEM of NcL.Fen.5B,85
Nd,Fe77.5BI8.5 mell-spunalloysmell-spunal l(e) m/s.
melt-spun alloys melt-spun at Vs =3(a), 5(b), 7(e)

Attempts to obtain information of the cooling behavior of a melt-spun


ribbon has shown progress recently by utilizing a digital thermograph recording
96 Satoshi Hirosawa

and analysis technique (Hirosawa et a!.


al. , 2001). Figure 4.9
4. 9 shows the time
dependence of the temperature of a melt-spun ribbon of Ndu45 Fen 8 18
18 ..55
= 3, 5 and 7 mis, which was obtained from the distribution of
quenched at V s =
temperature of the ribbons being cooled in the melt-spinning process captured
by the thermograph system. To convert spatial distribution of the ribbon
temperature into the time dependence, the distance along the ribbon from the
paddle position (Fig. 4. 9) was divided by the moving velocity of the ribbon,
which is assumed to be equal to the wheel surface velocity Vs (see the right
panel in Fig. 4.9).
4. 9).
The cooling rate for each wheel velocity is not a linear function of time as
is shown in Fig. 4. 9. Particularly, it becomes suddenly slow when the ribbon
departs from the wheel as is most evident for V s = 10 m/s. Average cooling
rates from the initial temperature of the melt, which is about 1520 K, to the
time when the ribbon departs from the wheel may be roughly estimated to be
4. 3, 2.
4.3, 7, 1.9
2.7, 1. 9 and O. 84 in the unit of 10 5
0.84 KI s, respectively, for V s = 10, 7, 5
and 3 m/s. The above observations suggest that the critical cooling rate Rc
corresponding to the nose of CCT
eeT curves for crystallization of y-Fe, Nd2 Fe14 8,
and Fe38 is about, respectively, 0.84, 1.9 and 4.3 ml
m/s.
s.

Temperature at the paddle

12S0 'c
c.a., 1250 'C

~\//0.84 105 K/s for VVs=3 m/s


\\/0.8410
!\\;:; \ /1.9 105 K/s for Vs=S
Vs=5 m/s
\\V
/I{ .2.7 105 K/s for Vs=7 m/s
\04.3
\(;;4.3 10 105 K/s for V
Vs=10 m/s
=10
900 :.:~:
900
". ).
:,: :
8S0 ::' .;,
850 11: -3m/s
800 11:: \ \.
.......... 5 m/s
-"-'S Nozzle
f.J
f..J 750
7S0 \ '1a.
.... -7m/s
~~ 7 00,
700),. ~"""" -lOm/s
-IOm/s
e
~ 550
650
~ 600
500
'\. '\
"\.

~~
\.
....-.........
................

SOO I-----'--'<--~-----------I
r----'---'<--~----------i
..............
...............
""II-
"'"II--

-.
Paddle
-
Vs
V,

d
Ribbon

..
4S0 1 - -
450 -- -- -- ---- -- -- -- -- -- -----------1
\
Quenching wheel
400 0 10 20 30
Time (ms)

Figure 4.9 Continuous cooling behavior in the melt-spinning process of NdNd4.5


u Fen 8 ,85
,8 . 5 as

evaluated by the infrared thermograph system for different wheel surface velocities.
Processing and Properties of Nanocomposite Nd,
Nd2 Fe" B-Based . . . 97

4.4 Transformation of Amorphous or Nanostructured


Precursors

As discussed in Section 4. 2, crystallization behavior of amorphous Nd-Fe-B


alloys to produce nanocomposite permanent magnets may be discussed in
terms of CCT diagrams. However, there exists little information of
crystallization kinetics of these materials to construct reliable CCT diagrams
for crystallization reaction. For NduFenBl85'
Ndu Fen B 185 , a model CCT diagram may be
proposed even though it is only a vague image of the whole story.
The typical composition range of the Fe3 B/Nd2Fel4 Fe14 B nanocomposites is
3 at. % -5
- 5 at. % Nd and 18at. % -19at.
- 19at. % B, the balance being Fe (Fig. 4.10).
4. 10) .
When Nd concentration exceeds this range, formation of Nd2Fe23B3 prevails.
This phase decomposes to Fe and Nd 1 . I Fe4 B4 in the Nd-rich alloys, excluding
l .1Fe4

the Nd2Fel4
Fe14 B formation, in the ternary alloys. Therefore, hard magnetic
properties are not obtained when Nd concentration exceeds the above range.
\
\
\

\
\ I
I NdllFe4B4
~ -------_\:;~
\ I
\ I \
/ \ I \
\ I \ I
c\f" \ \ I
~. \, I '\ II
~ \ ------- /\--------~'k-----
o~c/ Fe)B \ \ I
/ "\ I
/ "\
~ \!
\ /
\;' \ I
\, \
--~---------j--------~-
I \ I \ 1\
t\lloy composition /"
\ I \ I \
\ I \ I \
I \ I \ I \

I 2Nd Fe
B
23 3 " / "
:\-------7'-------~\--------:\

\\ I /
/
,\
\
/\
/\
Nd 2Fe\ l4I B3 \
\1
/
/\
/" \
\ I
/
/
;\
\
"
\

Fe 20 40
~
R.(at.%) . -.

Figure 4. 10 Compositions of typical Fe3 B/Nd


BINd,2 Fe" B type nanocomposite magnets and
related phases in the Nd-Fe-B ternary alloys.

Crystallization of amorphous alloys for production of Fe3 B/Nd2Fel4


Fe14 B type
nanocomposite permanent magnets proceeds in two stages. The tetragonal
Fe3 B crystallizes first, and the second stage crystallization of residual
amorphous follows to form Nd2Fel4
Fe14 B. Typical differential scanning calorimetric
(DSC) patterns of Ndu u Fen B 18
I85.5 during a continuous heating are shown in
Fig.4.11.
98 Satoshi Hirosawa

5 K/min
10 K/min
20 K/min
40 K/min
[00 K/min
100

]~
J~
1\

700 750 800 850 900 950


Temperature (K)

Figure 4.11
4. 11 DSC curves for amorphous Ndu Fen
Fe,,8'B5
B'85 recorded at various heating rates.
The anomalies of base lines at around 770 K are associated with a sudden change of
heating rate (not intrinsic, see text).
text) .

Transformation in isothermal aging may be described on a time-


temperature-transformation (TTT) diagram. To construct a TTT diagram,
isothermal phase transformation behavior has to be investigated. A melt-spun
Ndu Fen B 8 ,8 . 5 amorphous alloy exhibits formation of various phases as time
elapses in isothermal annealing and are shown in Table 4. 1.

4.11
Table 4. Crystalline phases occurred in Ndu Fen
Fe,,8'B5
B'8.5 during isothermal annealing.

Time
Temperature
0.6 ks 10. 8 ks 21 .6 ks

600'C Fe3B,
Fe38, Fe Fe Fe
Nd,2Fe23
Nd Fe,3 B
83 Fe3 8
Fe3B Fe3 8
Fe3B
Nd,Fe14 8
Nd2Fe14 B Nd2Fe'4 B
Nd,Fe'4 8 8
Nd2Fe14 B
Nd,Fe'4
630 'C Fe3 B
8 Fe Fe
Nd2Fe
Nd,Fe'3 83
23 B3 Fe38
Fe3B Fe38
Fe3B
Fe '4 B
Nd,Fe'4
Nd2 8 Nd, 8
Nd2Fe '4 B
680 'C
680'C Fe Fe Fe
Fe3B
Fe38 Fe3B
Fe38 Fe3 8
Fe3B
Nd2Fe14 B
Nd,Fe'4 8 NdFe,B4
NdFe484 NdFe 48 4
NdFe4B,

The result shown above cleatly


clearly indicates that the Fe3 B/Nd
8/Nd2Fe14
Fe'4 B
8
combination is only metastable and that the thermodynamically stable phases
are NdFe4 B4 and ex-Fe. Interestingly, the metastable phase Nd2Fe23 B
NdFe484 8 3, which
Processing and Properties of Nanocomposite Nd, Fe" B-Based . . . 99

is a cubic ferromagnetic phase near and above room temperature, appears


almost simultaneously with Fe3 B.
8. Ohmori et al. suggested that this is the
primary phase on which Fe3 B nucleates with lattice coherency (Ohmori et al. ,
Fe38
1999) .
1999).
The model TTT diagram based on this information is shown in Fig. 4. 12.
4.12.

r T
---------------------------------_.
__________________________________ 0

1400 ------ ---


1300 ,, ,-
Fe

1200

Fe3B
gh 1100 Nd 2 Fe l4
'4 B
Vs=3 m/s
V
1000 Vs=5 m/s ; Nd 2Fe23 B3
;Nd
Vs=IO m/s
900

800 '---------'--.,------'--::------'----,------'-::------'-:----'-::------'--:----
'--------'---,------'--::-----'---,------'-::------L,-------'-::----'-c,-----_ _
10-3 10-2 10- 1 100 10 10 1' 102 103
t (s)

Figure 4. 12
4.12 Schematic time-temperature-transformation (TTT) diagram for
Ndu Fen B 18
,8 . 5 (thick lines). Note that the nose temperatures for each TTT curve

have not been determined. The continuous cooling curves for various wheel
velocity are also shown by thin solid lines.

4.5 Direct Preparation of Nanocomposite Permanent


Magnets

As discussed previously, by choosing the CCT path correctly, a


nanocomposite permanent magnet may be produced directly in a single rapid
solidification process. Sometimes, the best magnetic properties are obtained
by this direct processing route. Magnetic properties as permanent magnets of
the as-spun alloys strongly depend on the cool ing rate, as can be suggested by
the arguments in the previous section. Figure 4. 13 shows dependence of
remanence (B r)'r), intrinsic coercivity (H CJ)'
cJ), and the maximum energy product
(BH ) max of NduFenB18.5
NduFen818.5 alloys melt-spun at different substrate velocities
(Hirosawa and Kanekiyo, 1998a). The microstructure of these alloys are
shown in Fig. 4. 9. Good hard magnetic properties are obtained only at Vs =
100 Satoshi Hirosawa

5 m/s with which the Fe3Fe3B/Nd2FeI4B


B/Nd2 Fe14 B nanocomposite structure is obtained.
The alloys obtained with Vs == 3 m/s have negligible coercivity because of the
large grain sizes and existence of the large ex-Fe dendrites (See Fig. 4.9). On
the other hand, alloys quenched at larger substrate velocities (V;;::::
(V s ?:-77 m/s)
have only small coercivity because of unfavorably small grain size and absence
the hard magnetic phase.
300 2.0
Nd4 Fe 77S B I85
E 250 P=1.3 kPa
:;;:
~ 1.5
"'-"""
~

'J!
~ 200
~

r~
1:2:!
~
150 * (BH)max

~ ~~
x He.!
~
E 100
E
A B,
~ j 0.5
50
V

~.~~:t=::::::==:;===::L---.J0
o 5 10 15 20 25
Substrate velocity Vs (I /ms)

Figure 4. 13 Magnetic properties of Fe3 8/Nd


B/Ndz Fel4
Fe14 B
8 nanocomposite prepared by direct
quenching of NduFenBl85
NduFen8185 alloys.

Directly quenched Fe3 B/Nd2 2 Fe14 B magnets obtained by the single-roll


method are thin plates of thickness in the 150 - 300 IJm range. The thickness is
inversely proportional to the surface velocity of the quench roll. Utilization of
such thin plate magnets has not yet been developed mainly because of the
fragility of the material.
In the case of Fe/Nd2 2 Fe14 B nanocomposites, the best magnetic properties
are often obtained bY by' the direct quenching. Formation of uniformly fine
microstructure probably requires a very short incubation period for
crystall ization, which is difficult to be achieved by the process in which over-
crystallization,
quenched alloys are annealed for further crystallization. To slow down the
kinetics, alloying additional elements is frequently effective. This point will be
discussed in the following section.

4. 6 Effects of Additive Elements on Crystallization


Behavior

Alloying a small amount of IVB - VIB elements, such as Cr, Nb, and Zr, has a
profound influence on solidification and crystallization kinetics of the Nd-Fe-B
Processing and Properties of Nanocomposite Nd, Fe,. B-Based . . .
Nd2 Fe" 101

alloys. Copper, on the other hand, is known to form atomic clusters in an


early stage of crystallization of Fe-based amorphous alloys because of its
imiscibility in Fe.

Among various elements which have been tried to improve magnetic properties
of the Fe3 B/Nd2Fe14 B-based nanocomposites, Cr is the most important
element to enhance the intrinsic coercivity, H cJ (Hirosawa and Kanekiyo,
1996). Due to the strong affinity of Cr with B, Cr is enriched in Fe3 B upon
crystallization and stabilizes this phase (Sano et ai., aI., 1998). This leads to
formation of the Fe3 B/Nd2Fe14 B composite even in the concentration range
where formation of Nd Nd2Fe23B3
2Fe23 B3 prevails (namely for Nd>c. a. , 5 at. %) in the
ternary alloys. Accordingly, the addition of Cr helps to realize nanocomposites
with a larger volume fraction of Nd2Fel4 Fe14 Band, hence a large coercivity.
Suzuki et al. pointed out another view that Cr has a significant effect on
kinetics of phase formation and decomposition, and that it allows a reaction
path in which the Fe3 B/Nd2 Fe14 B composite is formed as a metastable
intermediate structure instead of the Fe3 B/Nd2 Fe23 B3 combination (Suzuki
et ai.,
aI., 1999).
Niobium, on the other hand, stabilizes Fe23 Bs6 , and retards decomposition
of Fe3 B/N~ Fe14 B composite. Fe23 F~3 B6s has spontaneous magnetization of
approximately 1.7 T at room temperature (Kneller and Hawig, 1991), which is
larger than that of Fe3 B (1. 6 T) and Nd2Fel4 Fe14 B (1. 6 T). Therefore, the
presence of this phase may be beneficial. According to Ping et al. , Fe23 B6 s
crystallizes from the residual amorphous phase nearly simultaneously with
Nd2Fe14 B at a slightly higher temperature than the crystallization temperature
of Fe3B (Ping, 1999b). The resultant microstructure is characterized by finely
divided crystalline phases between Fe3 Fe3B B crystallites. Phases which appeared
and disappeared during isothermal aging in Nd u Fe bal B 185 Nb o6 are described
NduFebaIB18.5Nbo.s
by Hirosawa et al. (Hirosawa et al. , 2000).
Copper also has a prominent effect on kinetics of Fe3 B crystallization.
4.14
Figure 4. 14 shows traces of isothermal calorimetric analysis on
NduFe7S7B185CU03'
Nd u Fe76 7B I85 Cuo 3 Separate measurements of powder X-ray (XRD) on
specimens which were annealed for periods corresponding to the completion of
the first isothermal DSC peaks at the same temperatures as in the isothermal
DSC runs confirmed that the reaction was indeed the crystallization of Fe3 B
followed by that of Nd2Fe14 B. Crystallization
Crystall ization of Nd2Fe14 B in the isothermal DSC
runs was observed only at 830 and 835 K in this observation. The incubation time for
N~.5Fe77B18.5' It is to be noted
Fe3B is significantly shortened in comparison to NduFenBI8.5'
that crystallization temperature of Nd2Fe14 B also is significantly lowered by the
small amount of Cu (Hirosawa et al. , 2000).
Three dimensional atom probe microanalysis (3D-APM) has revealed that
102 Satoshi Hirosawa

835 K

830 K
_~--'-- 825 K

o 10 20 30 40
Annealing time (min)

Figure 4.14
4. 14 Isothermal DSC patterns of Ndu45 Fe76 7CUO 7 CUO 3B 185
'8 5 aged at temperatures
( T x) of Fe3 B.
slightly below crystallization temperature (T

Cu-enriched clusters are formed upon annealing of an amorphous


NduFe76 sB 1S . S CU02 alloy well below the crystallization temperature of Fe3 B.
sB,S.5CUO.2
The number of these clusters is numerous (10 24 //m m3) (Ping et al.,
al. , 1999a). Nd
is also enriched in these clusters, leaving an Nd-depleted and B-enriched
region around them. This region provides a chemically favored crystallization
site for Fe3BCPing
Fe3 B( Ping et al. , 1999b).
Copper is eventually resolved in the Nd2Fe'4 Fe14 B, while Nb was found to
partition in Fe23 B6 in the fully crystallized magnet. In the early stage of
crystallization, Nb is rejected from the Fe3B particles. Therefore, Nb is
believed to retard grain growth of Fe3 B, refining the grain size.
Fe3B,
Zirconium is also effective in refining grain sizes in the Fe3 B/Nd2Fe'4 B
nanocomposites. The addition of O. 1- 0.5 atomic percent of Zr increases the
crystallization temperature, particularly of Nd Nd2Fe14B CMiyoshi et al. , 2000).
2Fe'4 B (Miyoshi
The effect is greater than that of Nb. According to a recent 3DAP analysis of
Zr-Cu-doped magnets, Zr is rejected from Fe3 B and thus suppresses grain
growth of this phase. Zr is concentrated along grain boundaries between Fe3 B
and Nd 2Fe14 B in the fully crystall
Nd2Fe'4B ized stage (Kajiwara,
crystallized CKaj iwara, 2001).
The rare earth elements which have larger single-ion anisotropy than Nd,
namely, Pr, Dy and Tb, can be used to improve magnetocrystalline anisotropy
of the R2Fe'4 B phase. From the microstructural point of view, the heavy rare
earth elements give rise to a significant increase in crystallization temperature
of the hard magnetic phase, leaving the crystallization temperature of Fe3 B
(Coehoorn and de Waard, 1990). This is unfavorable
almost unchanged CCoehoorn
because grain growth of Fe3 B may occur before crystall ization of the R2Fe'4 B
crystallization
phase takes place. The simultaneous addition of elements with counteracting
effects on crystallization
crystall ization behavior is sometimes effective to avoid such
Processing and Properties of Nanocomposite Nd, Fe" B-Based . . . 103

difficulties, as is demonstrated in the following section.


Finally, Ga is quite important in improving magnetic properties of the
Fe3 B/Nd2Fe14 B nanocomposite (Kanekiyo et al., 1993). One of the best
Fe3 B/Nd2Fel4
Fe14 B-type nanocomposites has been Nd35 u Dy,
DYI Fe73 C03Gal
Ga, B
B'8
18 5 with
(BH)max of 136 kJ/m 3
(BH)""", 3
,, B,
B r of 1.18 T, and H He.]
cJ of 390 kA/m (Hirosawa and

Kanekiyo, 1994). Mishra and Panchanathan studied Nd NduDY1Fe73C03GalB18.S


35 Dy, Fe73 C0 3Ga, B'8.5

by TEM-EDX (energy dispersive X-ray) and observed that Ga additives were


all detected in the Nd2Fe'4 Fe14 B phase with Ga concentrations mostly near grain
boundaries (Mishra and Panchanthan, 1994). A fluorescence X-ray absorption
fine structure (XAFS) study using GaK-o: GaK-a edge indicated that homogeneously
distributed Ga in Ndu Fe73 Gal C03B B'85
I8 5 in the as-spun amorphous phase is
redistributed by being excluded from Fe3 B at the first crystallization stage.
After crystallization of Nd2Fe14 Fel4 B at the second crystallization stage, Ga was
suggested to form CsCI-type CoGa (Matsuura et al. , 1995). More recently, a
3D-APM analysis revealed that Co and Ga atoms are partitioned to the
Nd2Fe'4
Fe14 B phase, and evidence for a slight sl ight enrichment of Ga atoms at the
Nd2Fel4 B/Fe3 B interface has been found (Ping et al. , 1998). Effects of Ga on
Nd2FeI4B/Fe3B
the solidification and crystallization kinetics are to be investigated in more
detail.

4.6.2 Advanced Nanocomposite Permanent Magnets Based on


Fe3 B/Nd2 Fe 14 B
Fe3B/Nd2Fe14B
A major shortcoming of the Fe3 B/Nd2Fe'4 Fe14 B nanocomposite is its low
coercivity. In order to obtain a higher coercivity, there are basically two
factors to be improved. The first is the fraction of the hard magnetic phase,
Nd2Fel4
Fe14 B. A larger fraction of the hard magnetic phase (Nd 2Fe14 B) generates
a higher coercivity (Schrefl et al., 1994). However, when the Nd content
exceeds about 4.5 atomic percent, unfavorable soft magnetic phase
Nd2Fe23 B3 prevails, and destroys the hard magnetic properties in the ternary
Fe3 B/Nd2Fe'4
Fe14 B.
The second factor is the first magnetocrystalline anisotropy constant K, K 1 of
the hard magnetic phase. K 1 I of the Nd2Fe'4
Fe14 B-type compounds can be
improved by alloying heavy rare earth elements such as Tb and Dy. However,
grain size must be suitably refined in order to maintain the ratio between
exchange coupling and magnetocrystalline anisotropy energy in the adequate
range given by theoretical calculation (Fukunaga et al. , 1996).
Compositional adjustment has yielded a relatively high-coercivity material
on the basis of Nd-Fe-B-Cr-Co, namely Ndu4. 5Fen
Fe73 B
B'8.
18 .5Cr2
5Cr2 CO2 which has Br of
1.05 T, H kA!m, and (BH)max of 108 kJ/m 3
He.]cJ of 378 kA/m, 3
.. The simultaneous
addition of small amounts of Cu and Nb to this material has been successfully
carried out to improve (BH ) max with a slight sacrifice in H He.]
cJ to yield

NduFe723B,85Cr2C02CuQ2Nbo5 with B r of 1.10


NduFe723B18SCr2C02CU02Nbo5 1. 10 T, H He.]
cJ of 336 kA/m, and
104 Satoshi Hirosawa

(BH)max of 123 kJ/m 3


CBH)max J.
To utilize heavy rare earth elements, manipulation of crystallization
Fe14 B, can be carried out by readjusting Cu
behavior, especially that of Ndz2Fe'4
content so that the effect on the crystall ization temperature of Nd2
crystallization zFe14 B
counteracts with that of the heavy rare earth elements. An example of such
compositional modifications to yield higher coercivity magnets is
Nd35Dy,Fe713BluCr24Co24Cuo.4Nbo.5
NdJ5DY1Fe71JB185Cr24Co24Cuo.4Nbo.5 which has B, B r of 0.93 T, H CJ of 468
kA/m, and (BH) max of 100 kJ/m
CBH)max 3
J (Miyoshi aI., 2000). Utilizing the
CMiyoshi et ai.,
stronger effect of Zr to prevent grain growth, better magnetic properties have
been obtained in Nd34DY1F~17B185Cr24CoZ4Cu04ZrOZ
NdJ4DY1F~17Bl85Cr24C024Cu04Zr02 B r of 0.97 T,
which has B,
HcJ of 465 kA!m,
H CBH)max of 105 kJ/m3J .. Figure 4. 15 shows the coercivity
kA/m, and (BH)max
vs. remanence map as a summary of the compositional investigation based on
the information of micro-alloying effects on the crystallization behavior.
1.1

Nd4SFe7JBI8.SCr2CoZ
Nd4SFe73Bl8SCr2Co2

Nd4 SFe bal B I8 SCr2CoZCuO 4Nbos


Nd4SFebaIBI8SCr2Co2CuOANbos
1.0 . . .. Nd Febal B
4 SFebal
Nd4S BI8 Cr24C02ACUOANbo
sCrz
18 S 4C024CU04Nbo s

Nd34DYI BbalBI8SCrHCo2
Nd34DYI BbaIBI8SCr2.4C02 44 CU04
O4 ZrO2

Nd34DYI BbalBI8 SCr244C0 24 Nd J sDYI FebalBI8SCr2


sCrz Feba1B18.SCr2 4COz 4-
4C024-
E 0.9 Nd3STbIFebaIBI8SCr24C02.4CU04Nbos
NdJsTbIFebaIBI8SCr24Co24CU04Nbos CUoANbos
CUo4Nbos
~
c:Q


NdssFe66B 18SCrSCos
NdssFe66BI 8SCrSCos

0.8

0.7 '--_--'-_ _-'-_ _--'-_ _-'--_ _'--_--'-_ _-'-_ _....J


0.7'-----------"------'--------'------'----'---------'------'------'
o 100 200 300 400 500 600 700 800
H eJ (kNm)
He.!

Figure 4. 15 A map of magnetic properties of the advanced nanocomposite permanent


magnets based on Fe3 B/Nd2 Fe" B.

ex-Fe/Nd
o:-Fe/Ndz2 Fe14 B-based nanocomposite is expected to have a higher saturation
magnetization, and therefore the potential to exceed single-phase Ndz2 Fe14 B
magnets.
magnets_ Fabrication of Ndz2 Fe14 B-based permanent magnets with trace of
o:-Fe was reported by Yajima et al. as early as 1988 (Yajima
ex-Fe al. , 1988).
CYajima et aI.,
Processing and Properties of Nanocomposite Nd2 Fe" B-Based . . . 105

lO Fe
Nd 10 Fesz Bes showed a distinct indication of coarse a-Fe precipitation in as-spun
ez B
ribbon. Partial replacement of Fe with Zr resulted in the significant
improvement of magnetic properties. High concentration of Zr was detected in
a grain boundary phase at a junction of four grains where Nd was depleted in
melt-spun Nd9 Fen Fe72.55 COlO Zr25
Zru B6 . Manaf et a!.
al. investigated microstructure of
melt-spun Nd lO Fe FeS4
e4 B6 by transmission electron microscopy (TEM) and showed
existence of a nm sized a-Fe ex-Fe particle at a grain junction of a-Fe
ex-Fe grains (Manaf
a!., 1993). Ribbons with significant amounts of ex-Fe were prepared by
et aI.,
Withanawasam et al. (1994). A sample having 55 % NdzFe14 Band 45 % ex-Fe
had a high reduced remanence of 0.78, which was attributed to significant
exchange-coupIing among constituent phases.
In contrast to the Fe3 B/Ndz Fe14 Fel4 B nanocomposite alloys, the amorphous
formability of melt of the Fe-rich alloys for the ex-Fe/Nd zFel4 Fe14 B-based
nanocomposite is relatively poor. Fast quenching in the initial stage of rapid
solidification is essential to prevent the CCT curve from intersecting the nose
of y-Fe crystallization, otherwise development of coarse y-Fe dendritic
particles will results. Bauer et a!. al. reported magnetic properties of a series of
ex-Fe/NdzFe14 B nanocomposite permanent magnets prepared by melt-spinning
under reduced He atmosphere (Bauer et al., 1996). The use of He gas was
shown to be helpful to enhance heat transfer from the ribbon to ambient
environment just after the ribbon departed from the quenching roll surface while
the reduced gas pressure of the ambient atmosphere helps establishment of
good thermal contact between the melt and the roll surface.
The effects of small amounts of additives such as Ti, V, Cr, Nb, Hf, Mo,
W, and so on on the solidification and crystallization kinetics in the
ex-Fe/NdzFe14 B-type nanocomposite alloys seem not different from those in the
Fe3B/Ndz
Fe3 B/Ndz Fe Fe14 B-based nanocomposite alloys. In most cases, good ex-Fe/Ndz
l 4B-based
Fe14 B nanocomposite magnets are obtained directly from the melt within a very
narrow range of cooling rate. Addition of refractory metals probably results in
slower crystalline growth of both Fe and Ndz Fe14 B particles during the rapid
solidification
sol idification process. Nb is one of the most effective elements in this sense
(Hadjipanayis et al. a!. , 1995).
When amorphous phase is successfully prepared, crystallization crystall ization
transformation takes place in two or three phases, depending on composition,
in the Fe-rich alloys for ex-Fe/NdzFe14 B nanocomposite magnets, with the
primary crystalline phase being ex-Fe (Withanawasam et al., a!., 1995). In some
alloys, the metastable intermediate phase of the TbCu7 structure forms before
Nd zFe14 B does.
NdzFe14B
Modification of growth kinetics in rapid solidification process by TiC
addition to Nd-Fe-B alloys was successfully applied to develop alloy
composition suitable for inert gas atomization process (Branagan et al. ,
1996). Kramer et al. a!. discussed the effect of TiC addition on the sol solidification
idification
process in NdzFe14 B-based alloys in terms of velocity of sol idification front and
solidification
temperature there relative to the peritectic temperature of NdzFe14 B formation
106 Satoshi Hirosawa

(Kramer
CKramer et al., 1997). According to their model, if the growth rate of
NdzFe14 B is significantly reduced by TiC addition, transition of the growth
crystall ization of 2-14-1 in Iiquid
pattern from rapid crystallization liquid supercooled under the
peritectic temperature to dendritic growth of Fe in liquid heated up above
peritectic temperature due to recalescence is suppressed. The dedritic Fe will
be replaced by peritectic formation of 2-14-1 as temperature of the Iiquid liquid is
cooled down rapidly and results in regions of coarse 2-14-1 grains. Growth
kinetics in undercooled Ndz Fe14 Fel4 B alloys with C and Ti or Mo addition were
recently studied by Hermann and Bacher) using the electromagnetic levitation
CHermann and Bacher, 2000). The growth velocity of the NdzFel4 B
technique (Hermann
phase was estimated to be from 1. 1 to 6.4 6. 4 mm/s, depending on the degree of
super cool ing, in NdzFe 14B
Fe14 B melt, whereas it was reduced to O. 3 - 2 . 5 mm/s in
2.5
(NdzFe14 B)094 CTiC)003 alloy. The considerable slowing down of the growth
CNdzFe14B)094
kinetics resulted in refinement of grain sizes.
Effects of Cr, Ti Ti,, Nb, Zr, Hf, Ta, and W on microstructure and
magnetic properties of nanocomposites composed mainly of ex-Fe and
NdzFe14 B phase with minor amount of ferromagnetic boride such as
NdzFe23 B3 have been studied by Chang et al. (1999). C 1999). The formation of
metastable NdzFe23FeZ3 B3 was found to be suppressed by addition of Cr, Ti, Nb,
and V in (Nd
CNdo 95 Lao 05) 9 5Fe78 MzB 10 lo .5' Thermal magnetic analysis indicated the
existence of ex-Fe and a Ndz Fe14 Fel4 B-type phase. Considerable refinement of
grain sizes was observed by TEM in (Nd Fe78 Crz B Io
CNdo 95 Lao 05 ) 95 Fen 10 .5 and in
(Nd
CNdo 95 Lao 05 )95 Fe78 Ti z B 105
)9 5Fe78 IO 5
.
. In a similar composition range, Chiriac et al.
reported that relatively good hard magnetic properties can be obtained by a
chill disk melt spinning technique with a small surface velocity of only 3 m/s in
the as-cast state of Nd8Fen Fe73 C05HfzBIz (Chiriac
CChiriac et al., 2000). Thermal
magnetic analysis indicated formation of NdzFel4B-type
NdzFe14 B-type phase and ex-(Fe,
ex-CFe, Co).

4. 7 Practical Properties of Nanocomposite Nd2Fe14B-


Nd1Fe14B-
Based Magnets

4. 7. 1
4.7. Magnetic Properties

Numerous papers have been published on nanocomposite permanent magnets


based on various combinations of soft/hard magnetic phases. The majority of
the works dealt with isotropic nanocomposites in which crystallographic texture
does not exist. Table 4. 2 shows a few examples of magnetic properties of
isotropic nanocomposite permanent magnets. These magnets have magnetic
properties comparable to commercially produced powders for resin-bonded
magnets, and therefore practical potential to be utilized in the same
Processing and Properties of Nanocomposite Nd, Fe" B-Based . . . 107

appl ication.

Table 4.2 Typical magnetic properties of isotropic nanocomposite permanent magnets. For
preparation methods, MS stands for melt-spinning, MA for mechanical alloying,
A for annealing, and N for nitrogenation.

Preparation B, H eJ ( BH)ma>
BH)m"
Composition Ref.
method (T) (kA!m) (kJ/m 3 )

Nd, Fe80 8
B,o
,O (1) MS-A . 20 191 93 .
Ndu Fe73 C0 3Ga, 8
C03Ga, ,85
B'85 (2) MS-A .21
21 340 128
Nd35 Dy, Fe73 C0 3Ga, 8
Fe73C03Ga, ,8 .5
B'85 (2) MS-A 1. 18 390 136
Nd5 5Fe66 Cr5 C058 '85
Nd55Fe66Cr5C05B,85 (3) MS-A 086 610 966
gFe85 B
Nd9 86 (4) MS 1. 10
1.10 485 158
Nd7Fe898,
Fe89 B, (5) MS 1..28
28 252 146
Nd8Fe875 8
Bu (6) MS 25 c a. 500
c. 185.2
Nd3 5Fegl
Nd3 5Fe9' Nb, 8
B335 (7) MS .45
1.45 215 115
gFen5
Nd9 Fen 5COlO Zr, 5B
COlO Zr'5 86 (8) MS 0.89 c. a. 640 130
(Ndo95Lao05),IFe665ColO
(Nd 095 Lao 05 )" Fe66 5COlO Ti,8,o
Ti, B,o 5 (9) MS-A 0.94 1282 146
Sm8 Zr3 Fe85
Fe8S Co, N,
Nx ((10)
10) MS-A-N 0.94 764 118
Sm'1.67 COS8.33 Fe30
Sm".67C058.33 ( 11) MA-A 0.97 600 101
References: (1) Coehoorn el
et al.
al.,, 1988a,
1988a; (2) Hirosawa el al. , 1993,
et aI., 1993; (3) Hirosawa el al. , 1998b,
et al., 1998b;
(4) Manaf elet al., 1993b,
1993b; (5) Inoue elet ai.,
al., 1995,
1995; (6) Bauer el
et ai., 1996, (7) Hadjipanayis el
aI., 1996; et ai.,
aI.,
1995; (8)Yajimaetal.,
(8)Yaiimaelal., 1988, Changel
1988; (9) Chang et ai., 2000, (10)
al., 2000; (10)Yoneyamaetal., 1995, (11) Majima
Yoneyama et al., 1995;
el
et al. , 1996

4.7.
4. 7. 2 Corrosion Behaviors
The absence of rare-earth-rich phases in the nanocomposite magnets with
lean rare earth content results in significant improvement of corrosion
behavior. Nd ,o lo Fe78 Zr, Bs8 in which grain boundaries are enriched in Fe and
depleted in Nd and Zr showed significantly smaller weight increase under the
hum id environment of 60'C,
humid 60 'c, 95%
95 % relative humidity
hum idity in comparison to
Nd lo Fe82
Nd,oFes2Bs B 8 (Yajima
(Vajima et al., 1988). Figure 4.16
4. 16 shows weight changes of
compression-molded resin-bonded magnets of Fe3 B/Nd2Fe" B-type
nanocomposites Ndu3 8Fe74 Fe7' B,s
B I8 sCro 3C0 3(A), Nd, s Fen
Fe73 B,s
B '8 . s Cr2 CO 2 (C)
( C) ,
and Nd s s Fe66 B ,s I8 . s Crs Cos (D), cx-Fe/Nd Fel4 B-type nanocomposite
o:-Fe/Nd2 Fe"
Nd8sFes6
Fe86 B6 , and conventional Nd2 Fe l ,4 B-type magnet Nd l2 '2 Fen Cos B6
'
((H irosawa et al., 1999).
Hirosawa 1999 ). The magnets were made from powder of about
150 IJm in diameter and 2 weight percent of CI-free epoxy-resin. The small
difference in oxidation rates amongh A, C, and D 0 specimens may be
attributed to the slight difference in the rare earth concentrations. Namely, an
108 Satoshi Hirosawa

alloy with more rare earth content oxidizes faster.

40

~ 30 o NdI2FenCosB6(MQP-B)
of: (; NdsFeS6B6(MQP-Q)
t. Nd s FeS6B6(MQP-Q)
CD
E
~ 20
x
~
"D

10

(11)
Time (h)

Figure 4. 16 Weight change per unit surface area (dW) of compression-molded resin-bonded
magnets made of Fe3 B/Ndz Fe" Fe14 B-type nanocomposites Nd38 Fe7,B, 85 CrQ3C0
Fe74 B 185 CrQ3 C03(A),
(A) ,
Ndu Fe73 B l85 Crz COz (C), and Nds,s
Nd45Fe73B,8SCr2C02(C), Nd55 Fe66 B'8S
B 185 Cr5
Crs C0
Cos5 (D), ex-Fe/Nd
o:-Fe/Nd2 Fe14 B-type
z Fe"
nanocomposite Nd8Fe86 B6 , and conventional Nd2 z Fe"
Fe14 B-type magnet Nd'2
Nd 1Z Fen C0Cos5 B6
(Hirosawa et al. , 1999).

4. 7. 3 Magnetizability
In general, the magnetizability of isotropic nanocrystalline permanent magnets
increases with increasing H HcJ ' The smaller H cJ of rare-earth lean magnets in
comparison to that of "single-phase" Ndz2 Fe'4
Fe14 B isotropic magnets is desirable
in applications in which the magnets are used unsaturated due to the lack of a
large enough magnetizing force. Such a situation is frequently encountered in
multi-pole rotors of, for instance, small stepping motors. The magnetization
process in nanocomposite permanent magnets starts with rotation of magnetic
moment in the soft magnetic phases that is followed by rotation of magnetic
moments in the hard magnetic phase. The process is nearly reversible when
the magnetizing force is small. Irreversible magnetization process occurs only
when the rotation of hard magnetic moments is irreversible. Therefore, the
initial magnetization curve is "s" shaped. It follows that a nanocomposite with
a large amount of soft magnetic phase tends to be more difficult to magnetize
than a nanocomposite with a small amount of soft magnetic phase if H CJ cJ values

are the same. Micromagnetic simulation of magnetizability of nanocomposite


Nd-Fe-B magnets from direct current (DC) and thermally demagnetized states
was compared by Schrefl and Fidler (1998). A thermally demagnetized state
stores a higher amount of eschange energy than the DC demagnetized state.
Processing and Properties of Nanocomposite Nd, Fe" B-Based . . . 109

Only a small external field is required to initiate the irreversible switching of


entire grains in the thermally demagnetized state, leading to a higher initial
susceptibility.
susceptibi Iity.
Obviously, however, a magnet with a small H CJ value is demagnetized by
a small demagnetizing force. There is a trade-off between magnetizability and
stability against demagnetizing force. Practically, the stability is strongly
connected with thermal losses.

4. 8 Application Examples

A few evaluation studies of device performance with nanocomposite permanent


magnets have been reported for motor applications by Yamashita et al.
(Yamashita et al., 1998; Yamashita and Yamagata 1999). Two rotor
configurations suitable for small DC motors with mechanical output of 1 to 100
W (Fig. 4.17) were studied. The inner permanent magnet (IPM) rotor was
found to be more efficient than the surface permanent magnet (SPM) rotor
because IPM utilizes the reluctance torque associated with magnetic flux path
in the rotor. The Fe3 B/Nd2 Fe14 B-type nanocomposite was found to be suitable
for injection-molding of the thermoplastic compound into the curved gaps,
which serve as permanent magnet layer after magnetizing. In this case, the
morphology of the Fe3 B/Nd2 Fe14 B-type nanocomposite powder particles,

Q-axis Q-axis

. I !PM

Figure 4. 17 Rotor configuration of typical magnet motors. The parts 1 and 2 are,
respectively, permanent magnet and laminated steel core (Yamashita and Yamagata
1999) .
110 Satoshi Hirosawa

which is more round in comparison to thin flake-like morphology of


conventional melt-spun Nd-Fe-B, was beneficial to achieve low viscosity of the
compound in the injection process. The particular morphology of
Fe3 B/Nd2Fe14 B-type nanocomposite is closely related to the fact that preferred
Fe)
roll velocity is much slower than the conventional Nd-Fe-B in the rapid
solidification
sol idification process, a result of the large amorphous formability
formab iIity of the alloy.
Application of Nd-Iean magnets with relatively small H cJ values for multi-
pole rotors for small stepping motors may also be proposed. The obvious
difficulty is to maintain thermal stability
stabil ity in a practically tolerable range.
Dependence of irreversible flux losses on H cJ values of several Nd-Fe-B-based
nanocomposite was studied also by Yamashita et al. (2000). At least
400 kA/m coercivity seemed to be required to suppress the irreversible losses
in motors.
Apart from motors, various applications
appl ications may be possible for Nd-Fe-B
nanocomposite permanent magnets making use of their excellent corrosion
resistance, high remanence, and potentially low material cost. The
application area may include hybridizing this class of hard magnetic powder
with other powders of conventional Nd-Fe-B or hard ferrite. The low rare
earths content is certainly one major advantage of this class of materials
especially when a large amount of usage is expected, for instance, in
automobile applications.

4.9 Future Subjects

The processing of isotropic nanocomposite permanent magnets has been


discussed in this article. To realize the giant energy product which is predicted
for nanocomposite magnets, preparation of textured nanocomposites is
(Skomski and Coey, 1993) . Attempts to obtain textured
necessary (8komski
nanocomposite permanent magnets have been only partly successful, in that
textured multilayered composites have been fabricated with somewhat
disappointing values of remanent magnetization. The strong tendency of thin
films to develop anisotropy has been the only successful way to fabricate an
anisotropic nanocomposite. For instance, Parhofer et al. (Parhofer et al. ,
1996) prepared a sandwich Si02/Nd2FeI4B/cx-Fe/Nd2FeI4B/Cr
8i02/Nd2FeI4B/cx-Fe/Nd2FeI4B/Cr film by sputter
deposition with B, = 1.2 T and H cJ = 380 kA/m. More recently, Fullerton
et al. discussed magnetic proeprties of Sm-Co-based
8m-Co-based hard/soft magnetic
heterostructures and suggested that for 8m-Co/Fe
Sm-Co/Fe bilayers with suitably thin
constituent layers, (BH)
(BH ) max can be greater than that of Nd-Fe-B (Fullerton
et al. , 1999). The study of exchange-spring thin film magnets is conducted in
connection with exchange biasing, domain-wall magnetoresistance and domain
wall junctions.
Processing and Properties of Nanocomposite Ndz2 Fe" B-Based . . . 111

The way to generate texture in bulk nanocomposite permanent magnets


has not been discovered so far. Hot-pressed bulk nanocomposite magnets on
the basis of a-Fe/Pr2 Fe14 B have been fabricated using high pressure of 1 - 7
GPa (Wang et al., 2000). It was reported that increasing compression
pressure from 125 MPa to 5 GPa led to marked grain refinement in the magnet
and consequently resulted in significant improvement of magnetic properties.
Under high pressure greater than 5 GPa, crystallization was constrained and
the hot pressed magnet retained amorphous phase besides the Pr2 Fe14 Band
a-Fe. Further die-upsetting process to generate texture, which is known to be
effective for rapidly sol
solidified
idified and hot pressed Nd-Fe-B with excess Nd (Lee
et al. , 1985) has not been reported for hot-pressed a-Fe/R 2Fe14 B (R = Pr or
Nd). The solution-precipitation creep mechanism which assumes material
transfer via grain boundary Nd-rich liquid phase in the hot-deformation process
of fine crystall ine Nd-Fe-B alloys (Grunberger, 1998) may not apply in the
crystalline
a-Fe/R 2Fe14 B nanocomposites because there exists no grain boundary liquid
phase. Therefore, a completely novel process may be necessary to generate
texture in bulk nanocomposite magnets. Realization of textured
nanocomposites is certainly one of most challenging subjects for researchers in
this field.

References
Bauer, J.,
J. , M. Seeger and H. Kronmuller. J. Appl. Phys. 80: 1667 (1996)
Benjamin, J. S. Sci. Am. 234: 40 (1976) .
Branagan, D. J. , T. A. Hyde, C. H. Sellers and L. H. Lewis. IEEE Trans.
Magn. 32: 5097 (1996)
Buschow, K. H. J. , D. B. de Mo
Moijij and R. Coehoorn. J. Less-Common Met.
145: 601 (1988)
Buschow, K. H. J. In: K. H. J. Buschow ed. Handbook of Magnetic
Materials. Elsevier Science B. V. , p. 557 (1997)
Chang, W. C.,C. , S. H. Wang, S. J. Chang, M. Y. Tsai and B. M. Ma. IEEE
Trans. Magn. 35: 3265 (1999)
Chang, W. C. , S. H. Wang, S. J. Chang and Q. Chen. IEEE Trans. Magn.
36: 3312 (2000)
Chiriac, M., M. Marinescu and F. J. Castano. J. Appl. Phys. 87: 5338
(2000)
Chen, H. S. and C. E. Miller. Rev. Sci. Instr. 41: 1237 (1970)
Coehoorn, R., D. B. de Mooij, J. P. W. B. Duchateau and K. H. J.
Buschow. J. de Phys. Colloque C 8: 49 669 (1988a)
Coehoorn, R. , D. B. de Mooij and C. de Waard. J. Magn. Magn. Magn.
Mater. 80: 101 (1988b)
Coehoorn, R. and C. de Waard. J. Magn. Magn. Mater. 83: 228 (1990)
Davies, H. A. In: F. E. Luborsky ed. Amorphous Metallic Alloys..,
Butterworths, London, p. 8 (1983)
112 Satoshi Hirosawa

Duwez, P.,P. , R. H. Williams and W. Klement, Jr. J. Appl. Phys. 31: 1136
(1960)
Fukunaga, H.,H. , N. KitajimaandY.
Kitajima and Y. Kanai. Mater. Trans. JIM37:
JIM 37: 864 (1996)
Fullerton, EE. , J. S. Jiang and S.D.
S. D. J. Bader. Magn. Magn. Mater. 2000:
392 (1999)
Grunberger, W. In: L. Schultz and K. -H. Muller ed. Proc. 15th Int.
Workshop on Rare-Earth Magnets and their Applications. Werkstoff-
Informationsgesellschaft mbH, Hamburger, 333 (1998)
Hadijipanayis, G. C., L. Withanawasam and R. F. Krause. IEEE Trans
Magn. 31: 3596 (1995)
Hermann, R. and I. Bacher. J. Magn. Magn. Mater. 213: 82 (2000)
Hirosawa, S., H. Kanekiyo, H, M. J. Uehara. Appl. Phys. 73: 6488
(1993)
( 1993)
Hirosawa, S. and H. Kanekiyo. Trans. Mat. Res. Soc. Jpn. 14 B: 969
((1994)
1994)
Hirosawa, S. and H. Kanekiyo. Mater. Sci. Eng. A 217-218: 367 (1996)
Hirosawa, S. and H. Kanekiyo. In: L. Schultz and K.-H. Muller ed. Proc.
15th Int. Workshop on Rare Earth Magnets and Their Applications.
Werkstoff-Informationsgesellschaft mbH, Frankfurt, Germany, p. 215
(1998a)
Hirosawa, S. H. Kanekiyo and M. Uehara. J. Magn. Soc. Jpn. 22, Suppl. S
1 : 325 (1998b)
1:325
Hirosawa, S. , H. Kanekiyo, Y. Shigemoto. Mat. Res. Soc. Sypmp. Proc.
Vol. 577: p. 141 (Materials Research Society) (1999)
Hirosawa, S.,S. , Y. Shigemoto, K. Murakami and H. Kanekiyo. In: Proc. of
the 11 th International Symp. on Magnetic Anisotropy and Coercivity of
Rare-Earth-Transition Metal Alloys. The Japan Institute of Metals, Sendai,
Japan, p. S127 (2000)
Hirosawa, S. , T. Miyoshi, H. Kanekiyo and Y. Shigemoto. Presented at the
8th Joint MMM-Intermag Conf. (200 1) and to be published in IEEE Magn.
(2001)and
Inoue, A. and T. Masumoto. In: Y. Sakurai, Y. Hamakawa, T. Masumoto,
K. Shirae and K. Suzuki, ed. Current Topics in Amorphous Materials:
Physics and Technology. Elesevier Science Publishers B. V., p. 177
(1993)
Inoue, A.
A.,, A. Takeuchi, A. Makino and T. Masumoto. IEEE Trans. Magn.
31: 3626 (1995)
Kajiwara, K., K. Hono and S. Hirosawa. Mater. Tran. JIM, 42: 1858
(2001)
Kanekiyo, H., M. Uehara and S. Hirosawa. IEEE Trans. Magn. 29: 2863
(1993)
Kanekiyo, H. and S. Hirosawa. J. Appl. Phys. 83: 6265 (1998)
Kaneko, Y. In: Proc. 16th Int. Workshop on Rare Earth Magnets and Their
Applications. Japan Institute of Metals, Sendai, Japan (2000)
Kneller, E. F. and R. Hawig. IEEE Trans. Magn. 27: 3588 (1991)
Processing and Properties of Nanocomposite Nd,
Nd2 Fe" B-Based . . . 113

Kramer, M. J.,J. , C. P. Li, K. W. Dennis, R. W. McCallum, C. H. Sellers,


D. J. Branagan and J. E. Shield. J. Appl. Phys. 81: 4459 (1997)
Lee, R. W. , F. G. Brewer, N. A. Schaffel. IEEE Trans. Magn. 21: 1958
( 1985)
Liebermann, H. H. and C. D. Graham, Jr. IEEE Trans. Magn. 12: 921
(1976)
Majima, K. , M. Itoh, S. Umemoto, S. Katuyama, H. J. Nagai. Magn. Soc.
Jpn. 21: 629 (1996) (text in Japanese)
Manaf, A.A.,, M. AI-Khafaji, P. Z. Zhang, H. A. Davies, R. A. Buckley, W.
M. Raniforth. J. Magn. Magn. Mater. 128: 307 (1993a)
Manaf, A. , R. A. Buckley and H. A. Davies. Magn. Magn. Mater, 128: 302
(1993b)
Matsuura, M., S. H. Kim, M. Sakurai, K. Suzuki, H. Kanekiyo and S.
Hirosawa. Physica B 208 &. & 209: 360 (1995)
Mishra, R. K. and V. Panchanathan. J. Appl. Phys. 75: 6652 (1994)
Miyoshi, T., H. Kanekiyo and S. Hirosawa. In: Proc. of the 16th
International Workshop on Rare Earth Magnets and Their Applications. The
Japan Institute of Metals, Sendai, Japan, p. 495 (2000)
Ohomori, Y. , Y. Kadoya, K. Nakai, S. Hirosawa and H. Kanekiyo. In: M.
Koiwa, K. Otsuka and T. Miyazaki ed. Proc. of the International Conf. on
Solid-Solid Phase Transformations '99, The Japan Institute of Metals,
1999, Sendai,
Senda i, Japan
Parhofer, S. M., J. Wecker, C. Kuhrt, G. Gieres and L. Schultz. IEEE
Trans. Magn. 32: 4437 (1996)
Ping, D. H.,
H. , K. Hono and S. Hirosawa. J. Appl. Phys. 83: 7769 (1998)
Ping, D. H. , K. Hono, H. Kanekiyo and S. Hirosawa. J. Magn. Soc. Japan
23: 1101 (1999a)
Ping, D. H., K. Hono, H. KanekiyoandS.
Kanekiyo and S. Hirosawa. Acta Mater. 47: 4641
((1999b)
1999b)
Sakurada, S. , A. Tsutai, T. Hirai, Y. Yanagawa, M. Sahashi, S. Abe and
T. Kaneko. J. Appl. Phys. 79: 4611 - 4613 (1996)
Sano, N. , T. Tomida, S. Hirosawa, M. Uehara and H. Kanekiyo. Mater.
Sci. Eng. A 250, 146 (1998)
Schrefl, T. and J. Fidler. J. Appl. Phys. 83: 6262 (1998)
Schrefl, T. , J. Fidler and H. Kronmuller. Phys. Rev. B 49: 6100 (1994)
Senno, H. and Y. Tawara. IEEE Trans. Magn. 10: 313 (1974)
Skomski, R. and J. M. D. Coey. Phys. Rev. B 48: 15,812 (1993)
Suzuki, K. , J. M. Cadogan, M. Uehara, S. Hirosawa and H. Kanekiyo. J.
Appl. Phys. 85: 5914 (1999)
Umeda, T. , T. Okane, W. Kurz. Acta Mater. 44: 420 (1996)
Wang, Z. C. , S. Z. Zhou, Y. Qiao, M. C. Zhang and R. Wang. J. Magn.
Magn. Mater. 218: 72 - 80 (2000a)
Wang, Z., Zhou S., Zhang, M. and Y. Qiao. J. Appl. Phys. 88: 591
(2000b)
114 Satoshi Hirosawa

Withanawasam, L, G. C. Hadjipanayis and R. F., J. Krause. Appl. Phys.


75: 6646 (1994)
Withanasawam, L. L.,, A. S. Murthy, G. C. Hadjipanayis, K. R. Lawless and
R. F. Krause. J. Magn. Magn. Mater. 140: 1057 (1995)
Wecker, J. and S. Schultz. J. Appl. Phys. 62: 990 (1987)
Wecker, J. , K. H. Schnitzke, H. Cerva and W. Grogger. Appl. Phys. Lett.
67: 563 (1995)
Yajima, K. , H. Nakamura, O. Kohomoto and T. Yoneyama. J. Appl. Phys.
64: 5528 (1988)
Yamashita, F.,
F. , S. Hashimoto. Matsushita Technical Journal. 44: 190 (1998)
Yamashita, F. and Y. Yamagata. J. Magn. Soc. Jpn. 23: 1117 (1999) (in
Japanese)
Yamashita, F., F. K. Ohara, Y. Yamagata and H. Fukunaga. J. Magn.
Soc. Jpn. 24: 431 (2000)
Yoneyama, T. , T. Yamamoto and T. Hidaka. Appl. Phys. Lett. 67: 3197
(1995)

The body of experimental data presented in this article was obtained by the author's
author' s
colleagues at Sumitomo Special Metals. Co. Ltd., H. Kanekiyo, Y. Shigemoto, K.
Murakami, T. Miyoshi, and Y. Shioya.
5 Amorphous and Nanocrystalline Soft Magnetic
Materials: Tailoring of Magnetic Properties,
Magnetoelastic and Transport Properties

Arcady Zhukov and Jul ian Gonzalez

s. 1 Introduction

It is now nearly 30 years since the first metallic glass was produced by rapid
quenching from the liquid state by Miroshnitchenko and Salli (Miroshnichenko
and Salli, 1959) and Duwez et al. (Duwez et al., aI., 1966; Klement et al. ,
1970). This proved to be a crucial point in opening up new fields of research in
material science, magnetism and technology, such as metastable crystalline
phases and structures, extended solid solubilities of solutes with associated
improvements of mechanical and physical properties, nanocrystalline,
nanocrystall ine,
nanocomposite and amorphous materials which, in some cases, have unique
combinations of properties (magnetic, mechanical, corrosion, etc.).
Technological development of the fabrication technique and studies of the
structure, glass formation ability
abil ity and thermodynamics and magnetism of
amorphous alloys were intensively performed in 1960s - 1970s. Main attention
was paid to amorphous ribbons. These aspects have been extensively
analyzed in few review papers and books (Duwez, 1966; Jones, 1973;
Luborsky, 1983).
Most commercial and technological interest has been paid to soft
amorphous and nanocrystalline magnetic materials. Initially it was believed
that ferromagnetism could not exist in amorphous solids because of lack of
atomic ordering. Gubanov (Gubanov, 1960) theoretically predicted in 1960
that amorphous solids would be ferromagnetic. Later it was found that the 3d-
metal based amorphous alloys obtained by rapid-quenching of the melt are
excellent soft magnetic materials, i. e., they exhibit very low value of the
coercive field and relatively high saturation magnetization (Luborsky, 1983).
Such magnetic softness originates from the absence of magnetocrystalline
anisotropy in these alloys (Luborsky, 1983). The amorphous ribbons obtained
by the melt-spinning technique have widely been introduced as the soft
magnetic materials in 70th years. Their excellent magnetic softness and high
wear and corrosion resistance made them very attractive in recording head and
microtransformer industries.
116 Arcady Zhukov and Julian Gonzalez

Concerning the magnetic order in materials having structural disorder


(such as is the case of amorphous and nanocrystalline alloys), there are some
fundamental questions related to the existence of such well-defined magnetic
order. In fact, considering the ferromagnetic interactions of the magnetic
materials, it can be immediately considered a ferromagnetic structure. In this
naive idea the magnetic anisotropy effects have been neglected. Magnetic
moments tend to arrange their orientations parallel to each other via exchange
interactions; this they do when lying along a magnetic easy axis which is in the
same direction at every point in the material. However, if the easy axis
orientation fluctuates from site to site, a confl ict between ferromagnetic
coupl ing and anisotropy arises. As long as we imagine lattice periodicity, a
coupling
ferromagnetic structure is a consequence of ferromagnetic exchange
interactions, the strength of the anisotropy being irrelevant. In this situation
we are assuming a major simpl
simplification,
ification, namely, the direction of the easy axis
is uniform throughout the sample. With this simple picture we present crucial
questions related to the influence of an amorphous structure on magnetic orqer.
Regarding the magnetic order in amorphous and nanocrystalline
materials, we know that it stems from two contributions: exchange and local
anisotropy. The exchange arises from the electron-electron correlations. The
mechanism of the electrostatic interactions between electrons has no relation
to structural order and is sensitive only to overlapping of the electron wave
functions. Magnetic anisotropy is also originated by the interaction of the local
electrical field with spin orientation, through the spin-orbit coupling.
Therefore, magnetic anisotropy is also a local concept. Nevertheless, the
structural configuration of magnetic solids exerts an important influence on the
macroscopic manifestation of the local anisotropy. As a consequence, when
the local axes fluctuate in orientation owing to the structural fluctuation
(amorphous and nanocrystalline materials as examples), calculations of the
resultant macroscopic anisotropy become quite difficult.
In the case of amorphous ferromagnetic alloys, the usual approach to the
atomic structure of a magnetic order connected to a lattice periodicity is not
applicable. These materials can be defined as solids in which the orientation of
local symmetry axes fluctuate with a typical correlation length I = loA. The
local structure can be characterized by a few local configurations with
icosahedral, octahedral, and trigonal symmetry. These structure units have
randomly distributed orientation. The local magnetic anisotropy would be
larger in these units with lower symmetry. In general, are characterized by
fluctuations of the orientation local axis. It is remarkable that in these types of
structures the correlation length, I, of such fluctuation is typically the
correlation length of the structure and ranges from loA (amorphous) to 10 nm
(nanocrystals) and 1 mm (polycrystals).
Fluctuations in the interatomic distances associated with the amorphous
structure should also contribute to some degree of randomness in the magnetic
interactions of the magnetic moments. Nevertheless, such randomness is
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 117

expected not to affect the magnetic behavior qualitatively


qual itatively (Luborsky, 1983;
Handrich, 1969). Moreover, random distribution of the orientation of the easy
axis drastically affects the magnetic properties. A random anisotropy model
developed by R. Alben et al. (Alben et al., 1978) provides a successful
explanation about how the correlation length, I, exerts a relevant influence on
magnetic structure. The important question is: What is the range of
orientational correlation of the spins? Let L be the correlation length of the
magnetic structure. If we assume L I, the number of oriented easy axes in a
volume L 3 should be N = = (( L /I II)) 3 .
The effective anisotropy can be written as:
K eff = KIN 1/ 2 (5.1)

where K is the local anisotropy whose strength is assumed to be uniform


everywhere. By minimizing the total energy with respect to L, the following
expression can be deduced:
(5.2)

where A is the exchange stiffness parameter.


Special attention has been paid, in the last decade, to the study of
nanocrystalline phases obtained by suitable annealing of amorphous metallic
ribbons owing to their attractive properties as soft magnetic materials
(Yoshizawa et ai., al., 1988; Herzer, 1990; Herzer, 1995; Hernando and
Vazquez, 1993; Hernando et ai., 1998; Gonzalez et ai., 1997; Hernando et
al.,
ai., 1995; Murillo and Gonzalez, 2000). Such soft magnetic character is
though to be originated because the magnetocrystalline anisotropy vanishes the
very small magnetostriction value when the grain size approaches 10 nm
(Yoshizawa et al., 1988; Herzer, 1990a; Herzer, 1995) . As was
theoretically estimated by Herzer (Herzer, 1990a; Herzer, 1995), average
a-Fee Si) grains is negligibly small when grain
anisotropy for randomly oriented ex-Fe(SD
diameter does not exceed about 10 nm. Thus, the resulting magnetic behavior
can be well described with the random anisotropy model (Herzer, 1990;
Herzer, 1995; Hernando and Vazquez, 1993; Hernando et ai., 1998;
Gonzalez et al., 1997). According to this model, the very low values of
coercivity in the nanocrystalline state are ascribed to small effective magnetic
anisotropy (K eff around 10 J/m 3 ). However, previous results (Hernando
et al. , 1998; Gonzalez et ai.,
al., 1996) as well as those recently publ ished by
Varga et al. (Varga et al. , 2000) on the reduction of the magnetic anisotropy
from the values in the amorphous precursors (-- 1000 J/m 3 ) down to that
obtained in the nanocrystalline alloys (around 300 - 500 J/m 3 ) are not
sufficient to account for the reduction of the coercive field accompanying the
nanocrystall
nanocrystallization.
ization. The enhancement of the soft magnetic properties should
therefore be linked to the decrease of the microstructure-magnetization
interactions. These interactions, originating in large units of coupled magnetic
moments, suggest a relevant role of the magnetostatic interactions, as well as
118 Arcady Zhukov and Julian Gonzalez

in the formation of these coupled units (Hernando et al., 1998; Gonzalez


ai., 1996). In addition to the suppressed magnetocrystalline
et aI., magnetocrystall ine anisotropy,
low magnetostriction values provide the basis for the superior soft magnetic
properties observed in particular compositions. Low values of the saturation
magnetostriction are essential to avoid magnetoelastic anisotropies arising
from internal or external mechanical stresses. The increase of initial
permeability with the formation of the nanocrystalline state is closely related to
a simultaneous decrease of the saturation magnetostriction. Partial
crystallization of amorphous alloys leads to an increase of the frequency range
where the permeability presents high values (Mohri, 1984). These high values
in the highest possible frequency range are desirable in many technological
applications involving the use of ac fields.
It is remarkable that a number of workers have investigated the effects on
the magnetic properties of the substitution of additional alloying elements for
Fe in the Fe73.5 CUl
CUI Nb 3 Si 13 . 5B9 alloy composition so-called Finemet, to further
improve the properties, e. g., Co (Marin et ai., aI., 1997), AI (Lim et al. ,
1993; Tate et ai.,
al. ,1998)
1998) varying the degree of success. Moreover, it was
shown in Frost et al. (1999) that substitution of Fe by AI in the classical
Finemet composition led to a significant decrease in the minimum coercivity,
H~in~o. 5 Aim, achieved after partial devitrification, although the effective
magnetic anisotropy field was quite large (around 7 Oe) (Aranda et al. ,
2001). The coercivity behavior was correlated with the mean grain size and a
theoretical low effective magnetic anisotropy field of the nanocrystalline
samples was assumed in contradiction with those experimentally found in
Gonzalez et al. (1996); Varga et al. (2000) and Aranda et ai., al. , (2001).
Although amorphous Fe-, Co- and Ni-based ribbons are slightly more
expansive compared with conventional soft magnetic materials, such as
sendast, ferrites and superalloys (mostly due to the significant content of Co
and Ni), they found considerable applications in transformers (400 Hz), ac
powder distributors (50 Hz), magnetic recording as a magnetic heads and
mostly in magnetic sensors. The main reason for using amorphous alloys as
soft magnetic materials is a saving of the electric energy wasted by magnetic
cores. Besides, the combination of high magnetic permeability and good
mechanical properties of amorphous alloys may be successfully used in
magnetic shielding and in magnetic heads (Mohri, 1984). Thus, production of
about 3 millions heads per year in Japan in the mid-1980s has been reported
(Mohri, 1984).
The internal stresses result as the main source of magnetic anisotropy in
amorphous and nanocrystalline materials due to the magnetoelastic coupling
between magnetization and internal stresses through magnetostriction.
Consequently, these materials result to be very interesting for field and stress-
sensing elements because the Fe-rich amorphous alloys eXhibiting exhibiting high
magnetostriction values (A (Ass = 5 ) and therefore many of magnetic
= 10- 5)
parameters (i. e., magnetic susceptibility, coercive field, etc.) are
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 119

extremely sensitive to the appl ied stresses.


applied
According to Mohri (1984), magnetic sensors containing amorphous and
nanocrystalline materials are classified into three main groups:
1. Magnetometers using almost zero magnetostrictive amorphous and
nanocrystalline materials (flux sensors, current sensors, proximity sensors,
and magnetometers) .
2. Stress sensors using high magnetostriction constant (A s = 2 x 10 - 5 - 4 . 5 x
10 - 5).
5 ). These include:

(1) Sensors using bistable materials (pulse generators, encoders and


proximity sensors)
(2)
( 2) Stress-magnetic sensors (torque sensors, force sensors, knock
sensors)
(3) Ultrasonic sensors (distance sensors, frost sensors, pressure
sensors) .
3. Thermal sensors utilizing the Curie temperature (T = - 50 - 250 C ) .
Recently, the giant magneto-impedance (GMI) effect, which open new
possibilities for designing of quite sensitive magnetic field sensors (Panina and
Mohri, 1994; Beach and Berkowicz, 1994) was discovered.
This GMI effect consists of the large change of the electric impedance of a
magnetic conductor when it is subjected to an axial dc magnetic field. It has
been recognized that the large sensitivity of the total impedance of a soft
magnetic conductor at low magnetic fields and high frequencies of the driven ac
current originates from the dependence of the transverse magnetic
permeability upon the dc magnetic field and the skin effect. Besides, the
magnetoelectric effects (inverse Wiedemann and Matteucci effects) have been
widely used too. Therefore two novel groups of sensors have been introduced:
4. Magnetic field sensors based on the GMI effect of amorphous and
nanocrystalline materials.
5. Sensors based on magnetoelectric effects (inverse Wiedemann and
Matteucci) .
The main interest of the GMI effect is related to the high sensitivity of the
impedance to an applied magnetic field, achieving up to 500% relative change
of impedance in conventional amorphous wires with vanishing
magnetostriction. It should be outl ined that such GMI effect is generally about
outlined
less than 100 % in amorphous and nanocrystalline ribbons (Aragoneses et al. ,
2000) .
The amorphous wires, typically around 125 IJm in diameter, obtained by
the so-called in-rotating-water quenching technique, were first introduced a
few years ago. The magnetostrictive compositions exhibit rectangular
hysteresis loop, while best magnetic softness is observed for the nearly-zero
magnetostriction composition. Their main technological interest is related to
the magnetic softness in nearly-zero magnetostriction composition, magnetic
bistability in non-zero magnetostriction compositions and the GMI effect
(Vazquez et al. , 1997>.
1997). These magnetic properties result in a great interest in
120 Arcady Zhukov and Julian Gonzalez

sensor applications. Besides, the mechanism of the remagnetization process


changes with the sign of the magnetostriction constant, being such mechanism
as of nucleation of domain walls for Fe-rich composition and as of propagation
for Co-rich wires.
The latest advances in magnetic materials are based on the
miniaturization of modern magnetic materials. The alternative technology of
rapid quenching - the Taylor Ulitovski method to produce tiny metallic wires
(in the order of 3 to 30 IJm in diameter) covered by an insulating glass coating
has been recently employed for fabrication of ferromagnetic microwires coated
by glass (Vazquez and Zhukov, 1996a; Zhukov et ai., a!., 1999; Zhukov et a!.
al. ,
2000a) . It has been observed that Fe-rich compositions with positive
magnetostriction constant show generally rectangular hysteresis loop. On the
contrary, Co-rich negative magnetostrictive compositions have almost non-
hysteretic magnetization curves and glass coating removal results in the
appearance of magnetic bistability (Vazquez and Zhukov, 1996a; Zhukov
et a!.
al. , 2000a). The glass coating introduces additional internal stresses due
to the difference between the thermal expansion coefficients of glass coating
and metallic nucleus. Therefore, microwires of the same composition can
show different magnetic properties because of the different magnetoelastic
energy. Depending on the thickness of the glass coating, the switching field
(applied magnetic field necessary to observe magnetic bistability) is generally
one order of magnitude higher than for melt-spun wires.
On the other hand, when the magnetostriction constant, As' is close to
zero, a great variety of magnetic effects can be observed, depending on the
sign of As. It has been found that the hysteresis loop changes from unhysteretic
for slightly negative magnetostriction constant to rectangular for positive
magnetostriction (Zhukov et a!.al. , 2000).
In the case of glass-coated microwires, the glass coating introduces an
additional magnetoelastic contribution to the magnetic anisotropy acting as a
new parameter determining the magnetization process. One of the assumptions
for the explanation of this effect was the change of the sign of the
magnetostriction constant under the effect of internal stresses.
Recent studies permit us to obtain semi-hard magnetic behavior by
adequate thermal treatment (Zhukov et a!. ai.,, 1999).

5.2 Magnetic Properties: Magnetization Process

There are few reviews dealing with fundamental magnetic properties of


amorphous materials (Graham and Egami, 1978; Guntherodt, 1978; Wolfarth,
1980). Since the main challenge of this chapter is to review technological
magnetic properties, these features will be mentioned briefly.
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 121

Variations of the average magnetic moment J..I IJ and Curie temperature T c


for amorphous alloys as a function of transition metal, T, content have been
studied by Mizoguchi et al. (1973). They studied the dependence of IJ J..I and T c
on the number of valence electrons for amorphous T ao B 10
Tao lO P IO
lO thin films. (Where

T is a transition metal) Differences observed for the amorphous and crystalline


alloys for the case of strong ferromagnets were attributed to a charge transfer
from the metalloid atoms to the 3D transition metal. In the case of Tao B20 20 and
Tao P20 20 amorphous alloys such charge transfer of 1. 6 and 2. 4 electrons per
boron and phosphorus atoms is suggested to explain the observed dependence
of IJ
J..I on average transition metal valence electron concentration for several
amorphous ribbons (O'Handley (0' Handley et al. , 1976).
On the other hand, various authors have explained the difference in the
temperature dependence of the magnetization in crystall crystalline ine and amorphous
alloys (Alben et al., 1978b). Thus, the fitting at low temperature T 312
(1 (0) = ( 1 -
(1 ( T) /I (J BT 312/ + ... ) suggested the
3 2
behavior according to the (J
existence of spin-wave excitations in amorphous alloys. A larger value of the
B coefficient observed for the amorphous alloys has been attributed to the
chemical disorder. Inelastic neutron scattering confirmed the existence of well-
defined spin-wave excitations (Gambino and Saran, 1979).
On the other hand, generally flatter magnetization curves obtained by
plotting (J (1 ( T) /I (1 ( 0) versus T
(J (0) TIT
/ T c of amorphous metaIImetallic ic alloys have been
attributed to exchange fluctuations (Handrich, 1969). . Such exchange
fluctuations were attributed mainly to the local disorder (Alben et al. , 1978b).

5.2.1 Amorphous Ribbons: Effect of the Composition and Applied


Stress
The hysteresis loop of melt-quenched amorphous ribbons, measured along the
ribbon axis, is fairly rectangular, exhibiting low coercivity of the order of 1-
10 Aim (Fujimori et al. , 1974). In addition, the remanence is relatively small
and the magnetization beyond the coercivity increases slowly so that saturation
is not observed until
unti I a field of the order of 1-
1- 10 kA/m.
kAI m. The magnetic domain
observation has revealed the existence of maze domains (Obi et ai., al. , 1976).
It was concluded that the maze domains are attributed to the perpendicular
magnetic anisotropy arising from the strong magnetoelastic anisotropy in such
direction (large value of the saturation magnetostriction). The effects of the
composition, heat treatment and applied stress on magnetic anisotropy have
been extensively studied (Takahashi et al. , 1977; Luborsky et al. , 1975). It
can be concluded that the origin of such perpendicular anisotropy is related to
magnetoelastic coupl ing. Particularly, the effect of heat treatment on
remanence to saturation, M,/ Mri M.,
M s' ratio has been studied in Fe 40 Ni 40
Fe40 40 P 14
I4 B6
122 Arcady Zhukov and Julian Gonzalez

amorphous ribbons (Luborsky et aI.,al . 1975). An Mr/M s increase with


annealing temperature, T ann , was found reaching almost 1 at T ann """ ~
350 - 400 'C. On the other hand,
hand. evolution of a stress relief. estimated by the
effect of the annealing on the curvature of the ring shaped ribbon, was
obtained. Strong correlation between evolution of M Mss and stress relief is
r/ M
Mrl

=
(defined as W = r
found. Alternatively, a direct correlation of the magnetization energy, W
f:a (Ms- M)dH),
s - with the compositional change of the

s' is observed (0' Handley. 1975).


saturation magnetostriction constant, i\As'
It was found also, that W decreases with the applied stress (Egami
and Flanders, 1976) and that the temperature dependence of W is
exactly the same, as the temperature dependence of i\As. s' It was found that
the magnetostriction constant of the amorphous ribbons of systems
(Col-xFex)75SilOBI5
(Col-xFex)7SSilOB1S and (Col-xFex)sOPI3C7
(Co1-xFex)SOP13C7 changes from positive to negative
values passing through zero in the vicinity of x = 0.094 (Fujimori et al. al .,
=
1975). Consequently, amorphous ribbons with x = 0.094 exhibit excellent
magnetic softness (Kikuchi et al.
al ., 1975).
It is well known that the relaxation of internal stress in amorphous
materials (before crystallization) produced by the thermal treatments is
connected with the magnetic softening (i. e. , decrease of the coercive field
with the thermal treatment)
treatment),, owing to microstructural changes (Kikuchi et al.
1975).
The shape of the hysteresis loop,
loop. in particular, its squareness character.
is determined by the magnitude and easy axis of the various magnetic
anisotropies presented in the material. In the as-quenched ribbons the
magneto-elastic anisotropy is considered to be dominant, depending on
fabrication parameters, such as the melt temperature.
temperature, wheel speed, wheel
temperature, sucking distance on the melt,
melt. metal purity, ribbon thickness and
crucible-to-wheel spacing as well as perturbations of these parameters with
the time (Luborsky and Liebermann,
Liebermann. 1981 a; Luborsky et aI., 1981 b; b;
Gonzalez, 1996). After annealing.
Gonzalez. anneal ing, the magnetoelastic anisotropy drastically
decreases and the remaining anisotropy should be ascribed to the directional
ordering, magnetic field or stress induced anisotropy (Gonzalez, 1996;
Nielsen and Nielsen, 1980). The shape of the hysteresis loop is then
determined by the resultant direction and magnitude of this directional order
anisotropy.
It was shown that the remagnetization process in ribbons takes place
mainly by displacements of one or two domain walls through the cross section
(0' Handley, 1976). In contrast.
contrast, the displacement of a single domain wall
along the axis from one end to the other takes place in magnetostrictive
amorphous wires giving rise to the appearance of magnetic bistability
(Humphrey et aI.,
al. 1987 a; Mohri et al.
al.., 1990).
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 123

5.2.2
5. 2. 2 Amorphous Wires: Large Barkhausen Effect

The study of magnetic properties of ferromagnetic wires is a relatively old


topic in technical magnetism due to their unusual magnetic properties. Thus, it
was shown by Sixtus and Tonks (1932) and in subsequent papers that quick
domain wall propagation related with the Large Barkhausen Jump could be
observed in positive magnetostrictive wires under tensile stress. Later special
thermal treatment under stress was designed by Wiegand (1981) in order to
obtain spontaneous magnetic bistability, which is very useful for various
sensors applications (magnetic encoding, position and rotatory sensors,
magnetic field sensors etc.).
etc. ). These Wiegand wires are made from CoVFe
alloy. This particular process permits us to obtain magnetically soft internal
axially magnetized core and relatively magnetically hard outer shell.
The so-called" In rotating water" technique to produce amorphous wires
was introduced in the 1980s. Such method essentially consists in rapid
quenching of metallic alloy in the rotating centrifuge containing water. Few
review papers regarding the fabrication method (Ogasawara and Ueno, 1995)
and magnetic properties (Vazquez and Hernando, 1996) of such ferromagnetic
wires have been published in the 1990s. The resulting amorphous wires have a
typical diameter of 120 IJm. It was found that ferromagnetic amorphous wires
obtained by such method exhibit peculiar re-entrant magnetization process
denominated as spontaneous magnetic bistability if the chemical composition
shows positive or negative magnetostriction constant. Such magnetic
bistability is characterized by a single and quite large Barkhausen jump (at
fields around 10 Aim) between two stable remanent states, resulting in a
macroscopic squared hysteresis loop (Humphrey et al. , 1987a; Mohri et al. ,
1990) . In this way, three main groups of amorphous wires can be
distinguished:
(1) Fe-based wires with positive (of the order of 10- 5 ) magnetostriction
constant.
6
(2) Co-based wires with negative (of the order of 10-
10 -6)) magnetostriction

constant.
(3) Co-based wires with vanishing magnetostriction constant.
Typical compositions of all 3 groups of amorphous wires are presented in
Table 5. 1 (Mohri et ai.,
al., 1990). It is important to note that wires exhibiting
vanishing magnetostriction constant, i\ s' do not exhibit spontaneous magnetic
bistability. Typical hysteresis loops of all three compositions measured at
50 Hz are presented in Fig. 5. 1.
The origin of the rectangular hysteresis loop observed in magnetostrictive
wires is explained from their particular domain structure. Generally it is
considered that the domain structure in the remanent state consists of an inner
axially magnetized single domain and an outer multi-domain shell with
124 Arcady Zhukov and Julian Gonzalez

1.51---~==:;=====:;~
1.51---~==:;===::::=;~
1.0 A >0
?,;>O
5

0.5 I
-{).5
-0.5
o
---ri
-1.0
-1. 5 ':-:----:-::,__-'-::---::'.,..--':--~---:':,____:'_:_____:'
-1.5.':-::-----:c:----'-:----:::'::-....."..--=":------,,'=-----:'-:c------::'
-80 --{i0--{)O -40 60 80
0.6
0.4
0.4 A,=O
E 0.2
~ 0
~
:: -0.2
-{).2
-{).4
-0.4
-{).6
-0.6
~-=----'-:,__~--:-----=':----"::----:'::.....J
L-6--:-:-0-_-C-=-0--:'-::---':------:::'::------,,'40:---6-:':0,---'
--{)O -40 4 40 60
0.3
0.2
0.1
o
-{).l
-0.1
-0.2
-D.2
-{).3
-0.3
L-:-':-::-_--::'::---_--:-_--::'::---_:-':-::--'
'-:-':-::--~=-----:------:":---:-':-::---'

Figure 5.1 Hysteresis loops of Fe-rich 0.,>0),


0,>0), Co-rich 0.,<0)
(A,<O) and Co-fe-rich
Co-Fe-rich
0,"'='0)
o.s""'O) amorphous wires.

Table 5. 1 Compositions and magnetic properties of as-cast amorphous wires.


Composition M,CT)
MsCT) M,IM,
M,/M, W (Aim) As (XlO- 6)
A,(XlO-

8 12.58i 15
Fe72.5 B12.5Si'5 1.3 0498
0.498 12 25
COO.25
(Feo. 75 COO.25 )725 8 '2 5Si'5
)72.5 B'2 58i 15 1.19
1. 19 0.467 12.8 20
Fe O.6 Coo.,
((FeO.6 COO.4 ))72.5812.5 8i 15
72 5B'2.5 Si '5 1.18
1. 18 0.476
0476 10.4
104 16
(Feu COO.5)
(FeO.5 COO 58125 Si'5
.5) 72. 5B'2.5 8i '5 1.11 0.48 11.2 15
(Fe03 Co07
(FeQ3CoO B '2.5S i '5
)72.581258i,5
.7)72.5 0.87 0.474 9.6
96 8
(FeO.06 COO.9
94, ))72.5 B 12 .5Si
72. 58'25 8i 15
'5 0.81 0.649 Not bistable -0.
Co72. 58'25
C0725 i '5
8i,5
B'2.5S 0.64 0.313 6.4 -3

transverse easy magnetization axis (Humphrey et al. , 1987a). Such an outer


shell has radial easy axis in the case of Fe-rich wires with positive
magnetostriction constant and circumferential easy axis in the case of Co-rich
negatively magnetostrictive wires (Fig. 5. 2). The large Barkhausen jump
appears when, upon the application of the anti-parallel axial field, the
magnetization of the inner core reverses at a certain critical field. Since the
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 125

switching of the magnetization occurs at that critical field, it is denominated as


a switching field. The main feature of such bistable behavior is that it is
impossible to reach the demagnetized state during the re-entrant
remagnetization since the magnetization remains at one of the stable remanent
states: positive or negative. The volume of the internal axially magnetized
core can be approximately estimated from the relation between remanent,
M
M"r , and saturation, M s ' magnetization as:
(5.3)

(a)

(b)

5. 2 Schematic representation of domain structure of Fe-rich (;.,


Figure 5.2 0,>0)>0) (a) and Co-
rich (;',<0)
0,<0) (b) amorphous wires.

Such magnetic bistability is characterized also by a critical length. It was


shown that the magnetic bistability disappears when the sample length is
below some critical. It was shown also that this critical length correlates quite
well with the demagnetization factor (Zhukov et aI., 1995a). This inner core
is uniformly magnetized along the wire length except the wire ends. It was
found (Severino et aI., 1992; Reininger et aI., 1993; Chen et al., aI., 1993;
Zhukov et al. , 1995a; Zhukov et al., 1995b) that the nucleation of reversed
domains takes place near the wire ends. Besides, the magnetic bistability is
absent for Fe-rich wires shorter than some critical length, L cr (Zhukov et al. ,
1995a; Chen et al. , 1993). It was shown also that the critical length of Fe-rich
wires is strongly correlated with the wire diameter, as well as with the value
of the demagnetizing field. It is of the order of 70 mm for Fe-rich wires with
diameter 0 = 125 IJm decreasing up to 5 mm for 0 = 10 IJm (Zhukov et al. ,
1995a). For the case of Fe-rich wires this critical length is directly correlated
to the penetration depth of the reversed domains at the wire ends acting as the
nucleation sites. Regarding the critical length, it can be correlated well with
the demagnetizing factor (Zhukov et al.,aI., 1995a) indicating that the closure
domains penetrate from the wire ends inside the internal axially magnetized
core destroying the single domain structure.
Less attention was paid to the studies of Co-rich wires. It was found that
Co-rich wires (0 = 125 IJm) with a negative magnetostriction constant has a
magnetization profile similar to Fe-rich wires, but the critical length for
126 Arcady Zhukov and Julian Gonzalez

magnetic bistability is somehow lower than in the case of Fe-rich wires


(Zhukov et aI.,
al., 1995b). This rectangular hysteresis loop also disappears
when the magnetic field is below some critical value denominated as the
switching field (Humphrey et aI., al., 1987a; Mohri et ai.,
aI., 1990). The
rectangular hysteresis loop could be interpreted in terms of nucleation or
depinning of the reversed domains inside the internal single domain and the
consequent domain wall propagation (Zhukov et al., 1995a). The perfectly
rectangular shape of the hysteresis loop has been related to a very high
velocity of such domain wall propagation.
Drastic change of the magnetization at applied magnetic fields above the
switching field and sharp voltage pulses appearing in the pick-up coil during
the large Barkhausen jump and associated with fast domain wall propagation
promise to be very useful for different technological applications (Larin et al. ,
1996; Zhukov et al. , 2000c).
Domain wall propagation has been observed in different kinds of magnetic
materials with a rectangular hysteresis loop. Particularly, domain wall velocity
has been successfully measured in Ni single crystals, conventional amorphous
wires and recently in glass-coated microwires (Chen et al. , 1995; Riehemann
and Nembach, 1984; Zhukov, 2001).
The domain wall velocity has been measured by using an experimental
set-up, like in the case of the classical Sixtus-Tonks method (Sixtus and
Tonks, 1932). This set-up consists of a long solenoid creating a uniform
magnetic field, a short exciting coil placed quite close to the end of the
solenoid and two short pick-up coils, separated by a distance, I. The short
exciting coil creates an additional magnetic field in a local region close to the
wire end, generating in this way a reverse magnetic domain in this area. This
reversed domain grows and propagates along the entire wire under the effect
of the magnetic field created by the long solenoid. To obtain the temperature
dependence of the domain wall velocity, the experimental set-up has been
placed inside a SQUID magnetometer.
Figure 5.3 shows the field and temperature dependencies of the domain
wall velocity, v measured in amorphous glass-coated Fe65 B 15 Si 15 C5. The
measured values of v were between 35 and 65 m/s. m/ s. Weak but detectable
temperature dependence of the domain wall velocity has been observed in the
temperature range of 5 - 300 K. A slight increase of the domain wall velocity
is observed as measuring temperature increases. Observed domain wall
velocity of glass coated microwire results to be much higher than in Ni single
crystals (0. 000 :s;; vv:S;;
00011~ ~ 10 m/s) but somehow lower than in the conventional
amorphous FeSiB wires (150~ (150:S;; v~500 2000Ci Chen
v:S;;500 m/s) (Zhukov et al. , 2000c;
et aI.,
al., 1995). Like in the case of Ni single crystals, significant deviations
from the linear field dependence v (H) have been observed. Lower domain
wall velocity of glass coated microwires (ascompared with conventional
amorphous wires) could be related to strong and complex internal stresses
induced during the fabrication process by the difference in the thermal
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 127

expansion coefficients of the metallic nucleus and the glass coating


(R iehemann and Nembach, 1984). Such an assumption
(Riehemann assumption is supported by tak ing
taking
into account the decrease of the domain wall velocity observed in amorphous
FeSiB wires under external tensile stress (Riehemann
(R iehemann and Nembach, 1984)
1984)..

-5 K
-5 K 0
0 -40 K
-40K , '
"
60 ... -lOOK
-100 K 0
0 -150K
-150 K :'(;i,
" -250 K /,~::

55 ,
..",6:''W ,~
I' ,
," ,f': "

I
,r; o'iJi :'
I",
I ,II,
, ,",

,
.,' s:/~:-c.,: ," I
, I, I

.. ,0,~':'
" ,/,',','
JI" ~"J'l'.
45
,.
~

o,'QA-'Q"
, ,'_-0 '
" "',,

,d/,d,'
,','",'
,+
I "

40 ---'" ,+,

70 75 80 85 90 95
I-J (Aim)
H(A/m)

Figure S.5. 3 Velocity of domain wall propagation measured in amorphous glass-coated


Fe65 8 '5 Si ,5 C5 microwire at 5 - 300 K.

The values of the magnetic field given on the horizontal axis in Fig. 5.3
are lower than the switching field. This means that after overcoming this first
highest energy barrier corresponding to the domain wall nucleation, the
domain wall propagates in a field even lower then the switching field. The
magnetic field required to overcome this barrier is higher than the intrinsic
coercivity of the material. The lowest limits of the magnetic field presented in
Fig. 5. 3 (around 70 Aim) could indicate the existence of a magnetic field
limiting the domain wall propagation. Below this magnetic field, the
propagation of a domain wall is not observed. The critical magnetic field, Her'
actually can be obtained by the extrapolation of the curves presented in
Fig. 5.3 at v == O. This extrapolation gives values of Her 5 - 10 Aim which are
an order of magnitude lower than the switching field. Both the critical field,
Her and the minimum field at which the domain wall propagation is observed
seem to be almost insensible to the temperature. The Her can be interpreted
as the intrinsic coercivity of the material and characterizes the efficiency of the
obstacles for the domain wall propagation.
It is remarkable that the existence of two critical fields - for the nucleation
(switching field) and for the propagation of the domain wall (critical field) -
has similar features. With the classical case of the nucleation, when to start
the phase transition, a critical nucleus size should be overcome. In our case,
it was shown that the reversed domains already exist at the wire ends due to
128 Arcady Zhukov and Julian Gonzalez

the effect of the demagnetizing fields. So, one can assume, that the
appearance of the Large Barkhausen Jump should be related to the critical size
achieved by one of the reversed domains achieved under the effect of an
appl ied magnetic field (Ponomarev and Zhukov, 1984; Zhukov, 1993).

5.2.3 Nanocrystalline Materials: Coercivity


Such as has been mentioned, the discovery of Fe-rich nanocrystalline alloys
carried out by Yoshizawa et al. (1988) was really important, owing to the
outstanding soft magnetic character of such materials. Typical compositions of
the precursor amorphous alloys, which after partial devitrification reach the
nanostructure character with optimal properties, are FeSi and FeZr with small
amount of B to allow the amorphization process; smaller amounts of Cu, which
act as nucleation centers for crystallites; and Nb, which prevents grain
growth. This effect is provided by Zr in FeZr alloys. After the first step of
crystallization, FeSi or Fe crystallites are respectively finely dispersed in the
residual amorphous matrix. In a wide range of crystallized volume fraction, the
exchange correlation length of the matrix is larger than the average
intergranular distance, d, and the exchange correlation length of the grains is
larger than the grain size, D. Magnetic softness of Fe-rich nanocrystall ine
alloys is due to a second complementary reason: the opposite sign of the
magnetostriction constant of crystall ites and residual amorphous matrix, which
crystallites
allows reduction and compensation of the average magnetostriction.
Figure 5. 4 shows the thermal variation of the coercive field (He) in a
eTa-containing) amorphous alloy. This behavior is quite similar to that
Finemet-type (Ta-containing)
shown in the case of Nb-containing ones and particularly, evidences the occurrence
of a maximum in the coercivity linked to the onset of the nanocrystallization process
(Vazquez et aI., 1994; Muller
MUlier and Mattern, 1994).

10

1I L,----:-~-_=_!:::_-~,----:-!.-::----::-!:-!:'----::-:!
'-::-----:-:!-::----=-!:-=--~,.----:-!:::----::~----:;-:!
o 100 200 300 400 600
Tann CC )

Figure 5.4 Evolution of the coercive field, He' with the annealing temperature, T qnn'
onn , in
Fe73SCul Ta3Sil3S8g
Ta3Si13.sB9 amorphous alloy.
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 129

Considering the grain size 00 to be smaller than the exchange length, Lex'
Lex'
and the nanocrystals are fully coupled between them, the random anisotropy
model implies a dependence of the effective magnetic anisotropy <K > with the
sixth power of average grain size, O. The coercivity is understood as a
coherent rotation of the magnetic moments of each grain towards the effective
axis leading to the same dependence of the coercivity with the grain size
(Herzer, 1990a>:
1990a):
6
<K> Ki0 6 ( <K> Ki0
He = Pc T
---:J; = Pc J A 3
ss
= ~
)
(5.4)

kJI m3 is the magnetocrystall


where K 1 = 8 kJ/m ine anisotropy of the grains, A =
magnetocrystalline
11
Jim
10- 11 JI m is the exchange ferromagnetic constant, J ss = 1.2
1. 2 T is the saturation
magnetic polarization and Pc is a dimensionless pre-factor close to unity. The
predicted 0 6 dependence of the coercive field has been widely accepted to be
followed in a 0 range below Lex (around 30 to 40 nm) for nanocrystall ine
Fe-Si-B-M-Cu (M= WA yA to VIA metal) alloys. (Herzer, 1995a; Suzuki et al. ,
1998). A clear deviation from the predicted 0 6 law in the range below He =
1 Aim was reported by Hernando et al. (1998). Such deviation was ascribed
to the effects of induced anisotropy (e. g. , magnetoelastic and field induced
anisotropies) on the coercivity could be significant with respect those of the
random magneto-crystalline anisotropy. As a consequence, the data of He(O) He (0)
were fitted by assuming contribution from CD the spatial fluctuations of induced
anisotropies and ~ (2) <<K (i. e., a dependence He = J(a
Hc(i.
K u> to He J (0 2 +(bO)6)
+ (bo)6)
was found with 0a = 1 Aim representing the contribution originating from the
induced anisotropies) .
In order to investigate the effect of the grain size on coercivity, we have
obtained this dependence of He (0) (D) in alloys treated by Joule heating.
Experimental results on this dependence are shown in Fig. 5. 5 (Murillo and
Gonzalez, 2000). The fitting of this dependence appears to follow a
surprisingly rough dependence of the H ce OC0cc0 3 - 4 type (the best regression was
found fitting the 0 3 law).
law ). It must be noted that our data of He (0)( D) correspond
to a grain size variation between about 10 to 150 nm. As it is well known, an
analysis of this He (D)
(0) data in terms of the random anisotropy model is only
justified if the grain size is smaller than L ex and, hence, could be not
appl icable (in the framework of the random anisotropy model) to the range
applicable
grain size above L ex which results to be only two points of our data in Fig. 5.5
(Murillo and Gonzalez, 2000). These points should correspond to a magnetic
hardening due to the precipitation of the iron borides. In this case the random
anisotropy model should be applied by taking into account the volume fraction
and the different anisotropy of the iron borides. This indicates that < K K1 >
should vary as 0 3 , contrary to the theoretical 0 6 law. This indicates that He is
mainly governed by <K 1>' 1 >, which varies as 0
3
contrary to the theoretical 0 6
law. This contradiction on the He ( D) 0) law between the theory and the
experimental has been recently explained by Suzuki et al. (1998) considering
130 Arcady Zhukov and Julian Gonzalez

the presence of long-range uniaxial anisotropy, K u' which influences to the


exchange correlation value, length and yields an anisotropy average given by:

(5.5)

0.1 '-----c-----'------'-----'------'------'------'--....L--'---'-:,-----
la' 102
D(nm)

Figure 5.5 Dependence of the coercive field, He' with the average grain diameter, 0,
for the Fen5 Nb3Sil3.sBg
Fe73. S Cu, Nb 3Si'3.5 8 9 amorphous alloy.

The second part of Eq. (5. 7) corresponds to (K 1) 1 > and if K u is coherent

1> ,, this second part


in space or if its spatial fluctuations are negligible to (K 1)
ultimately determines the grain size influence on the coercivity. Such influence
changes from the 0 6 law to a 0 3 - 4 one when the coherent uniaxial anisotropies
dominate over the random magnetocrystalline anisotropy. An additional point
in order to justify Eq. (5.7) is connected with the fact that the results of Fig.
5.5 were obtained in samples treated by the Joule heating effecteffect. This kind of
annealing could induce some inhomogeneous magnetic anisotropy, which could
be responsible for this significant change of the grain size dependence of
coercive field. This argument is supported by the remanence ratio, which is
achieved by these samples of values around O. 50. As a consequence, it can
be assumed that the presence of more long-range uniaxial anisotropies larger
than the averaged magneto-crystalline anisotropy (K 1 ). >. It should also be
noted that Eq. (5.7) can, interestingly, account for the occurrence of dipolar
and deteriorated exchange intergrain interaction, and thus can be more
realistic than the simple anisotropy averaging since those features are involved
in the accomplishment of a nanocrystallization process.
An interesting study cauld be one related to the dependencies of the
magnetization in partially crystallized samples. Figure 5. 6 displays such
dependencies. In Table 5.2 we present the evolution with the annealing time
of the average grain size, 0, the Si atomic percentage diffused in the Fe
crystalline lattice and the crystalline phase percentage. It has been mentioned
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 131

that with samples annealed with short times, (0. 5 to 5 min) there was not evidence
of crystallization. In samples treated with long times, the Si content was of the order
of 20% at and slightly larger in the samples having a larger grain size.
1.5 , - - - - - - - - - - - - - - - - - - - - - - - - - ,
1.5

1.2

E 0.9
2:.~

~ 0.6

0.3
1211
12h

0 60

Figure 5. 6 Current density dependence of the saturation magnetization of the


Fe735 Cu, Nb38i'3.5 8 9 amorphous alloy treated at 40 A/mm 2 with 0, 0.5.
Fe73.sCuINb3Si13sB9 1. 2.
0.5, 1, 2,
60. 120,
5, 60, 120. 300 and 720 min.

Table 5. 2 phase.
Evolution with the annealing time of the percentage of crystalline phase,
percentage of Si
8i content inside bcc phase and average grain size of
Fe73.s
Fe73.5 CUI
Cu, Nb33Si 13 . sB
8i'3.5 9 amorphous alloy ribbon treated by current annealing at
89
40 A/mm2 .
Crystalline phase 8i content inside
Si Average grain size
T ann (min)
(% ) bcc phase (%(%)) (nm)
60 60 19 12
120 65 21 15
300 70 21 15
720 78 22 17

On the other hand, the current density dependence of the coercivity for
Finemet-type alloys results to be a very interesting study (illustrated in Fig.
5.7). Such dependencies exhibit a peak of coercivity in nanocrystallized
samples (treated at 60 - 720 min). This peak of coercivity occurs above the
Curie point of the residual amorphous phase. The intensity and the width of the
peak strongly depend on the anneal
annealing
ing current density. It must be mentioned
that current density above the Curie point of the amorphous matrix being
paramagnetic and its thickness is high enough to avoid exchange interactions
between the grains. The nanocrystall ine sample can be magnetically
nanocrystalline
considered as an assembly of isolated or weakly magnetostatic interactive
single domain particles.
The coercivity behavior can be interpreted in the framework of the two-
132 Arcady Zhukov and Julian Gonzalez

80
.'\
i \
I \
~h
60
\

E IhiI \'
~40 j
15,h,... ~
{\ \
u
I ii
0
::r::
::t: '.
I ,i' I'
\' '12 h
20r-_....... ....,j~~i '~\ \
>''It.,

o 10 20 30 40 50
J (A/mm 2)

Figure 5.7 Variations of the coercive field with the current density of Fen5
Fens Cu,
CUI Nb3
3 Si 13s
l35 8g
E3g
2
samples current annealed at 40 A/mm with 0, 0.5, 1.
O. 0,5. 1, 2.
2, 5.
5, 60,
60. 120,
120. 300 and 720 min,min.

phase model (Hernando et al. , 1995). At room temperature the system is soft
because the exchange between crystallites is large enough to make the
correlation length larger than both the intergranular distance and the grain
size. As the current density rises and approaches to the Curie point of the
residual amorphous matrix, the exchange constant decreases and some grains
start to be weakly coupled. The exchange correlation length decreases and the
crystallites progressively start to act as pinning centers.
It is interesting to note the differences of the Curie point with the annealing
time (Fig. 5. 6) corresponding either the remaining amorphous matrix as well
as to the nanocrystalline phase. This behavior should be explained taking into
account the compositional change of the amorphous matrix (with progressive
loss of Si and Fe with the annealing time) which can significantly change the
Curie point of this phase. Unavoidable mixing of atoms at the interface
nanocrystal-amorphous gives rise to the formation of thin layers of alloys with
unknown composition. This lack of knowledge about the nature of this interface
brings up an interesting question related to the coupling between two phases
with large interface area, as is the case of these soft magnetic nanocrystalline
Fe-base alloys.

5. 3 Processing: Induced Magnetic Anisotropy

As-prepared amorphous alloys have a non-equilibrium or metastable


microstructure, attributed to the lability of the amorphous state as well as to
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 133

the high quenching stresses. In the case of glass coated microwires with
composite (glass-metal) structure, such stresses have crucial role. Therefore
various technological procedures might significantly change their
microstructure and consequently the magnetic properties. The most common
technological procedure used for the tailoring of magnetic properties of
amorphous magnetic materials is thermal treatment. Application of magnetic
field and/or applied stress can strongly affect atomic order and consequently
magnetic properties of amorphous and nanocrystall
nanocrystalline
ine materials. Thermal
treatment under applied stress and/or stress can also induce macroscopic
magnetic anisotropies, described below.
Stress relaxation and local environment changes during thermal treatments
generally result in evolution of the local atomic structure towards more stable
atomic configuration. It is well established
establ ished that generally the relaxation of
internal stress in amorphous materials (before crystallization) produced by the
thermal treatments is connected with the magnetic softening (K (Kikuchi
ikuchi et al. ,
1975). On the other hand there are various factors affecting soft magnetic
behavior of amorphous materials. At least five pinning effects have been
identified and discussed by Kronmuller
KronmLilier (1981) as contributing to the total
coercivity:
( 1) Intrinsic fluctuations of exchange energies and local anisotropies
(10- 3 -1 me), HeW.
<10-
(2) Clusters and chemical short ordered regions 1 me), He(SO).
(3) Surface irregularities 5 me), He(surf).
He (surf).
(4) Relaxation effects due to local structural rearrangements (0. 1 -
10 me), He(rel).
( 5) Volume pinning of domain walls by defect structures in
(5)
magnetostrictive alloys (10 - 100 me), He (a) (C1) .
Within the framework of the statistical potential theory, the resultant total
coercivity was expressed as following (Kronmuller
(KronmLilier et al. , 1979):
He(total) = (a)2 + H e(surf)2 + He(SO) 2+ H eW 2Jl!2
= [H e(cr)2 J1!2 + He(rel).
(5.6)

In the case when the surface irregularities give largest contribution, the
various terms add linearly, i. e. :
= He(cr)
He(total) = He((J) + He(surf)
He (surf) + He(SO) + HeW + He(rel). (5.7)
A detailed analysis of each term is described in (Kronmuller et ai., 1979;
Kronmuller, 1981).
Kronmuller,1981).
On the other hand it was found that the shape of the hysteresis loop is
determined by the magnitude and easy axis of the various magnetic
anisotropies presented in the material (Luborsky and Liebermann, 1981 a;
Luborsky et al., 1981 b; Gonzalez, 1996). Thermal treatments affect this
magnetoelastic anisotropy. After annealing, the magnetoelastic anisotropy
drastically decreases and the remaining anisotropy should be ascribed to the
134 Arcady Zhukov and Julian Gonzalez

directional ordering, magnetic field or stress induced anisotropy. The shape of


the hysteresis loop is then determined by the resultant direction and magnitude
of this directional order anisotropy.
It was found that the magnetostriction constant of the amorphous ribbons
of systems (Col-xFex)75SilOB15
(Co,-xFex)75SiIOB,5 and (Col-xFex)sOPI3C7
(Co'-xFex)80P13C7 changes from positive
to negative values passing through zero in the vicinity of x = = 0.094 (Fujimori
et al., 1975). Consequently, amorphous ribbons with x = O. 094 exhibit
excellent magnetic softness (Kikuchi et al.al.,, 1975).
On the other hand it is well known that so-called "nanocrystalline
materials", that is, systems consisting of two phases (nanocrystalline grains
randomly distributed in a soft magnetic amorphous phase) can be obtained in
certain amorphous FeSiB-based alloys with small additions of Cu and Nb by the
devitrification of the conventional amorphous FeSiB-based alloys after
600 'c, 1 h (i. e. , at temperatures between
anneal ing in the range of 500 - 600C,
annealing
the first and second crystallization peaks) (Yoshsizawa et al., 1988). After
crystallization, such material consists of small (around 10 nm grain size)
nanocrystals embedded in the residual amorphous matrix. In addition, the
devitrification process of these amorphous alloys (with trademark finemet)
leads to the possibility of obtaining rather different microstructures depending
on the annealing parameters, as well as on the chemical composition. These
materials exhibit extremely high magnetic softness at certain conditions of
thermal treatment. In fact, the macroscopic magnetic anisotropy averages out
(when L D) and the domain wall can move without pinning. For some
particular compositions and volume crystallized fraction the average
magnetostriction constant vanishes and the magnetoelastic contributions to the
macroscopic anisotropy also become negligible.
Recently modified Taylor technology has been employed for the
fabrication of glass coated microwires in an amorphous state and with classical
"finemet" composition for the metallic nucleus (Vazquez and Zhukov, 1996;
Zhukov et al. , 2000b; Zhukova et al., 1999; Zhukova et al. , 2000). The
main characteristic feature of such tiny amorphous materials is the appearance
of strong internal stresses owing to difference of thermal expansion coefficients
of the metal and the glass. Consequently, the magnetic properties could be
improved by adequate treatment: heat treatment or chemical etching of the
glass coating or by the selection of adequate chemical composition of the
metallic nucleus.
Thus the magnetic properties of amorphous and nanocrystalline materials
are correlated with their microstructure, composition and internal stresses
originated from the fabrication method.
The aim of this section is to present a brief review of the results on
compositional dependence and processing of amorphous and nanocrystalline
materials in order to design their magnetic properties.
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 135

5.3.1 Amorphous Metallic Alloys


The magnetization characteristics of amorphous magnets (in particular, those
of metall ic glasses produced in a form of thin ribbons) can be modeled by
metallic
annealing under the influence of externally applied magnetic field or mechanical
stresses, in a similar way as the case of crystalline soft magnetic materials.
This modeling is of technological interest since it allows tailoring the
magnetization curves of a given material to the specific requirements, thus,
broadening the range of its application. In fact, the amorphous alloys,
although macroscopically isotropic in the as-quenched state, can exhibit
macroscopic magnetic anisotropy if, such as has been mentioned, they are
subjected to suitable annealing treatments at the presence of either a magnetic
field (field annealing, FA) or a mechanical stress (stress annealing, SA).
More recently stress + field annealing has been performed in several
amorphous alloys and it has been shown, first, that the stress + field induced
magnetic anisotropy cannot be supposed as a simple sum of the stress-induced
anisotropy and the field induced anisotropy. Second, it seemed, from earlier
works on stress + field annealing (Vazquez et al., 1987a; Vazquez et al. ,
1987b), that the direction of the stress + field induced anisotropy was fixed by
the direction of the magnetic field applied during the induction process and the
applied tensile stress may enhance this anisotropy. But this hypothesis was
found to be untrue with posterior investigations, which are presented here.
Figure 5. 8a and b show the experimental data reported by Gonzalez and
Kulakowski (1989a) on the annealing temperature dependencies of different
kinds of induced magnetic anisotropies (H Tr : transverse field annealing; H p :
longitudinal field annealing; (J: stress anneal
anneal ing; a: ing; H rT + a:
annealing; (J: transverse field +

stress anneal ing and H p + a:


annealing (J: longitudinal field + stress anneal ing) in two
annealing)
CoFe-based (Co-rich) amorphous alloy ribbons. These results on induced
anisotropies point out that the field-induced anisotropy has the same direction
as the magnetic field, which was applied during the annealing.
The behaviors of the stress and stress + transverse field induced
anisotropies are similar. The evolution of these two kinds of anisotropy as a
function of the annealing temperature is quite different with respect to field
induced anisotropy and they increase with the annealing temperature reaching
a maximum value, K max' at the temperature, T max' and then decrease. This
decrease 6f stress and stress + field induced anisotropies after maximum
value allows us to assume that their origin could be ascribed to the magnetic
ordering of pair atoms at high annealing temperatures probably acting on the
metallic glass at equilibrium, while at low temperatures other processes and
mechanisms are occurring during the induction process, such as the
development of a longitudinal plastic component (Gonzalez et ai., 1990a;
Nielsen et ai.,
al., 1985) and the change of the local short range order which is
136 Arcady Zhukov and Julian Gonzalez

600
(Coo.9sFeo.OS)80Si lOB 10
(Coo.9sFeo.os>SoSi loB 10

H.L
H.l
400 -
- -
-0 -0-0-_
0 -<J-0-_

'~ 200
~
4
:.::
~
TannCC)
oa 100 200 300

200

K//II
400 K

(a)

600

400
~

E
~ 200
~

ao 100 200300 400 ;';'


~ ~
~.,
Tann (oC ) _ ~,.
Tann("C)_
~HII
~H//
200
200 K
K// II

(b)

50 8 Magnetic anisotropies induced in two nearly-zero magnetostrictive


Figure 5. magnetostriclive Co-rich
amorphous alloys.

introduced by the stress during the thermal treatment. If the stress is absent,
the former mechanism seems to be very weak and, in this case, the field-
induced anisotropy is almost symmetric with respect to the change of the
direction of the annealing field. The direction of the stress + longitudinal
induced anisotropy strongly depends on the composition. This dependence is
presented in Fig. 5.
5.9,
9, where it can be seen that the increase of the relative
concentration of Fe and Si gives the decrease of the longitudinal anisotropy.
As can be seen from Fig. 5.8a, 5. 8a, b, filed annealing at elevated
temperature but below the Curie temperature induces a macroscopic magnetic
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 137

Kl-(J/m 3 )
Kl-O/m Kl- Kl-
(Co1_xFex)7SSi IS BIO
(Col_xFexbSiISBJO .: (Co09sFeooS)90-xSi.,BIO
(Coo.9sFeooS)90-xSixBIO .: (Co09sFeo oS)7s Si xB2S-x
(Coo9sFeo oshsSixB2s-x
400 400 0: ColOo_x(Si
C01oo_xCSi o 6B04)x 400 C07sSixB2S_x
0: C07sSixB2S-x
I
/+
./
/
0 10 I 20
/
/

,
I

o 20 25 I 30
15 x
\
\ I
, x 20 x

\\I'0'
, II

400 400 KillII '


I{
1/
1- 400
K II 0 K III;
I

Figure 5.9 Influence of the composition on the direction of the stress + longitudinal field
induced magnetic anisotropy in Co-Si-B and Co-Fe-Si-B (Co-rich) amorphous alloys.

anisotropy with the preferred axis determined by the direction of the


magnetization during the annealing. Moreover, field induced anisotropy
annealing
increases as the anneal ing temperature increases, similar to the magnetization
with the temperature, vanishing at Curie temperature. This behavior has been
observed in metallic glasses with different composition (Fe-rich) (Miyazaki
and Takahashi, 1978; Luborsky and Walker, 1977); CoFe-based (Co-rich)
(Nielsen and Nielsen, 1985; Gonzalez and Kulakowski, 1989a); CoNi-based
(8Iancoetal.,
(Blanco et al. , 1991) and Co-rich (Gonzalezetal.,
(Gonzalez et al. , 1990b). The microscopic
origin of this field-induced anisotropy has been successfully explained
according to directional ordering of atomic pairs mechanism developed by Neel
(Neel, 1954; Luborsky and Walker, 1977; Miyazaki and Takahashi, 1978).
This model predicts a dependence of the field-induced anisotropy with the
annealing temperature as:
(5.8)
where n is a constant, the value of which can be assumed to be equal to 2 if
the microscopic origin is the directional ordering of atomic pairs. Theoretical
predicted value of the index n was experimental found in FeNi-based
FeN i-based metallic
glasses (Neel, 1954). Nevertheless, deviations of such theoretical value have
been obtained in metallic glasses of composition (Co,- (Co 1- x Fe xx )78 SilO B
8 12 (Chen
and Tai, 1985; Chen, 1987) and (Co Fe )75Si15810
(Co'-xFex)75Si,5BIO
-
1 x x (Gonzalez et aI.,
1990a). In this case, an additional contribution coming from the single-ion
(initially n = 3) can be accounted, which allows us to conclude that depending
on the annealing temperature the weight of each two contributions could be
different according to the content of magnetic elements.
The stress and stress + field induced anisotropies have been a topic of
intensive research in the last two decades in amorphous Co, CoFe CoFe,, CoFeNi
and CoNi-based alloys (Nielsen and Nielsen, 1980; Hilzinger, 1982; Nielsen,
138 Arcady Zhukov and Julian Gonzalez

1983; Nielsen et alal. , 1985; Vazquez et al.,


aI., 1987a; Vazquez et al., 1987b;
Gonzalez et al., 1987; Gonzalez and Kulakowski, 1989a; Gonzalez et al. ,
1990b; Blanco et al al. , 1991). It is a remarkable difference with respect to the
field anneal ing, that the stress and stress + field anneal
annealing, annealing
ing in metall
metallic
ic glasses
could induce magnetic anisotropy even at annealing temperatures above the
Curie temperature. Moreover, these kinds of anisotropy can be induced in
zero-magnetostrictive amorphous alloys and its origin is, therefore, different
to that of field induced anisotropy. Hernando et al al. (1985), Blanco et al al.
( 1991) and Gonzalez and Blanco (1992) have proposed several contributions
to explain the microscopic origin of the stress and stress + field induced
anisotropies, namely: CD atomic pair ordering and (2) CZ) tetrahedral Bernal holes
(BTH)
( BTH) surrounded by (3Co-1 Fe, 1Co-3Fe ), (3Fe-1 Ni, 1Fe-3Ni) and
(3Co-1 Ni, Co-3Ni). These BTHs differ from each other in the local anisotropy
according to particular interactions. In fact, the compositional dependence of
the stress and stress + field induced anisotropies of several amorphous alloy
systems (Fig. 5.10) show two maximum values for the (Co 1 -- .x Fe. Fe x )-based
system for around x = = 0.75 (Co-rich) and x == 0.25 (Fe-rich) respectively, the
Co-rich maximum being larger than that of the Fe-rich maximum. This implies
that the 3Co-1 Fe BTHs could give a more efficient contribution to the induced
anisotropy. Moreover, the absence of these maximums for Ni-containing (Co
and Fe) ,systems or the lower intensity of those maximums in CoFeNi-based
systems suggest that the Ni atoms seems to "dilute" magnetically the
amorphous matrix.

~
0.5 'q, ))
, I
\ I
\ /
\ I
" I
""0" /

o 0.5 1.0
x

Figure 5. 10 Compositional dependence of the stress + field induced magnetic anisotropy


5.10
(. )(Co'-xFex)7SSi,sBIO;
)(Co'-xFex)7SSilSBlO; (0) [Col-x(Feo.sNio.s)x]7sSilsBlO;
[Co,-x(Feo.sNio.s)x]7sSi,sBlO; (x) (Co'-xNix)7sSi,sBlO'
(@) correspond to the stress-induced anisotropy in ( Co,- x Nix )7S
The values of () B,o
hs Si,s B IO
amorphous alloys.

As has been mentioned, it is interesting to remark that these stress and


Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 139

stress + field induced anisotropies can be induced by annealing even above


the Curie temperature. This characteristic suggests that the source of these
induced anisotropies are not related to magnetoelastic effects but rather to
transformations of the microstructure probably connected to "viscoelatic" flow.
By analogy with the creep strain, these induced anisotropies have been
divided in two contributions: the anelastic, K an' of recoverable character and
the plastic, K pi' (permanent) one being:

Kind = K an +
an p,
K pi (5.9)
where Kind is the stress or stress + field induced magnetic anisotropy. The
anelastic component, K an' results in a transverse easy axis and is reversible
while the plastic anisotropy, K pi' results in an axial easy axis and is
irreversible in the sense that it is mostly retained even if further treatments
without applied stress are performed. When the first stress or stress + field
annealing is carried out, K an predominates as its kinetics permits rapidly
induced anisotropy, K pi is induced more slowly as viscoelastic flow is being
generated. As further treatment is performed at the same annealing conditions
but without applied stress, the resultant induced anisotropy changes the
preferred direction toward that of the ribbon axis. After the second (stress
free) annealing, only plastic deformation and thereby plastic anisotropy
remains. The strong influence of preannealing on the stress induced anisotropy
is notable (Fig. 5. 11 ). In general, preannealing favors the anelastic
component to be outstanding.
Moreover, experiments reported in (Blanco et ai., al., 1991; Gonzalez
et ai., 1991) have shown that the plastic component is identical for stress and
stress + longitudinal field and stress + transverse field induced anisotropies
in CoFe-based amorphous alloys. Consequently, the stress and stress + field
induced anisotropies can be considerably increased by preannealing
treatments. This increase could be related to relaxation processes, which are
reversible in nature. This conclusion seems to rule out the free volume
elements origin of the anelastic component of these anisotropies proposed by
Argon and Kuo (1980).

5.3.2
5.3. 2 Nanocrystalline Alloys
The magnetization characteristics of Finemet-type nanocrystalline magnets
(FeCuNbSiB-alloy), similar to those of metallic glasses, can also be well
controlled by the magnetic anisotropy induced by field annealing
anneal ing (FA), stress
annealing (SA) and stress + field annealing (SFA). Magnetic field annealing
induces uniaxial anisotropy with the easy axis parallel to the direction of the
magnetic field applied during the heat treatment. The magnitude of the field-
induced anisotropy in soft nanocrystalline alloys depends upon the annealing
<that is, if the magnetic field is applied during the
conditions (that
140 Arcady Zhukov and Julian Gonzalez

400

0'-----_--'--
O'-----_--J- '--_ _--'-
-'-- -'-- __
250 300 350
T.'T1ann (oC ))
111 (OC

Figure 5.
50 11 Influence of the pre-annealing on the stress + field induced magnetic
anisotropy as a function of the annealing temperature in (Coo (COO 50 Fe025
FeQ.25 Ni o.25 h5 Si ,5
025 15 8 10
amorphous alloy. (x) stress + longitudinal field annealing (as-quenched, AQ); ( )
(AQ); (0) stress + longitudinal field annealing (pre-annealed, PA); ((lj)
stress annealing (Am; ((j)
stress + transverse field annealing (AQ); ((lj)
((j) stress + transverse field annealing (PA).

nanocrystallization process or firstly the sample is nanocrystallized and then


submitted to field annealing) (Herzer, 1995a; Herzer, 1995b) and on the alloy
composition (relative percentage content of Fe and metalloids) (Herzer,
1994a). Nevertheless, this field induced anisotropy is induced at temperature
range 300 - 600'C
600C (above the Curie temperature of the residual amorphous
matrix and below the Curie temperature of the ex-Fe (Si) grains). Thus, the
anisotropy induced during nanocrystallization should primary originate from the
bcc grains. The amorphous matrix has a rather inactive part, since its Curie
temperature is far below the typical field annealing temperature.
The evolutions of the different types of anisotropy induced in a typical
alloy susceptible to be nanocrystallized, as a function of current density
(thermal treatment carried out by current annealing technique under action of
stress and/or field) are shown in Fig. 5. 12. As can be seen, stress and stress
+ field induced anisotropies increase with the current density (temperature)
up to a maximum value at 45 A/mm 2 which may be related to a maximum of the
coercive field. The increase of induced magnetic anisotropy up to 45 A/mm 2
could be ascribed to an increase of the intensity of the interactions between the
metallic atoms and, consequently, an increase of the induced anisotropy could
be expected. This argument is linked to the internal stress relaxation produced
by thermal treatment in the metallic glasses. Similarly, field-induced
anisotropy decreases with the current density down to minimum value. Such a
minimum is observed again at around 45 A/mm 2 . For current densities higher
than that of the maximum, induced anisotropy monotonically decreases with
current annealing density.
Studies on the stress-induced anisotropy (Murillo et ai.,aI., 1993; Nielsen
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 141

300

200

100
---
~

!l;;;x
~
0

~ -100
E
:.::
SA
-200

--300
300 '~-~-'-----'-----'------'-----'-----'-----'
---_'---_'--_'--_'--_-'--_-'--_-'------l
34 36 38 40 42 44 46 48 50
2
Jann (A/mm )

Figure 5. 12 Maximum magnetic anisotropies versus current annealing intensity induced in


Nb 3 Si 15 .5B, samples.
Fe73.5CU, Nb3Si,5587

et al. , 1994; Herzer, 1994b; Kraus et ai., aI., 1996; Hofmann and Kronmuller,
1996; Lachowicz et aI.,al., 1997)
1997> indicate that behaviors resembling those in
metallic glasses can also be found in nanocrystalline magnets Finemet-type.
Although the occurrence of this effect has been well confirmed, nevertheless,
its origin seems to be not entirely interpreted up to the present. Herzer
( 1994b) proposed an explanation, claiming that this anisotropy is of a
magnetoelastic nature and is created in the nanocrystallites a-Feex-Fe (Si) grains
due to tensile back stresses exerted by the anelastically deformed residual
THe above conclusion seems to be highly probable because
amorphous matrix. Ttie
of a strong correlation between the stress-induced anisotropy and the
nanocrystall ites found by Herzer (1994b). However,
magnetostriction of the nanocrystallites
Hofmann and KronmUlier
KronmUller (1996) and Lachowicz et al. (1997) suggested an
alternative explanation of the origin of the considered anisotropy. They
adapted Neel's
Neel' s calculations of atomic pair directional ordering (Neel, 1954)
to the conditions of the investigated material, obtaining theoretical value of the
energy density of the stress-induced anisotropy of the same order of magnitude
as that observed experimentally. Consequently, besides the magnetoelastic
nanocrystall ites suggested by Herzer, the directional
interactions within the nanocrystallites
pair ordering mechanism in a-Fe
ex-Fe (Si) grains is also a very probable origin of
the stress-induced anisotropy in Finemet-type material.
The occurrence of dipolar and deteriorated exchange intergrain interaction
should also be considered to explain the origin of the stress-induced anisotropy
in the nanocrystalline alloys (Gonzalez, 1996; Hernando et ai., aI., 1998). This
leads to a more realistic situation than the simple anisotropy averaging, since
those features are involved in the accomplishment of a nanocrystallization
process. In this way, the procedure to obtain the weighted average anisotropy
nicely proposed by Alben et al. (1978a) strongly depends on the degree of
142 Arcady Zhukov and Julian Gonzalez

magnetic coupl ing. This stress anisotropy is induced, as has been noted
coupling.
previously inside the grains. The maximum value (around 1000 J/m 3 ) is
clearly lower than 8000 J/m 3 corresponding to the magnetocrystalline
anisotropy of the ex-Fe ( Si) grains and, therefore, the origin of the stress
a-Fe (Si)
anisotropy should be strongly connected to the internal stresses in the FeSi
nanocrystals. An interesting question should be that related to the coupl ing
between these two phases with large interface area such as is the case of
nanocrystall ine alloys. For this, a deep knowledge about the nature of
Fe-rich nanocrystalline
the interface results is to be determined. Unavoidable mixing of atoms of the
interface gives rise to the formation of thin layers of alloys of unknown
composition, which makes this study very complicated.

5.3.3 Induced Magnetic Anisotropy in Glass-Coated Amorphous


Microwires
Such as has been mentioned, the Taylor-Ulitovski method is an alternative
technology to produce amorphous materials, that is, tiny metallic wires (in the
order of 3 - 30 IJm
I-lm in diameter) covered by an insulating glass coating with
very promising magnetic, magnetoelastic and magnetotransport properties. It
was found that Co-rich microwires with negative magnetostriction constant
present an almost non-hysteretic loop with a relatively high magnetic
anisotropy field up to around 8000 Aim. In contrast, Fe-rich microwires of
positive magnetostriction show rectangular hysteresis loops with switching field
depending on diameter of metallic nucleus and thickness of glass coating
(Vazquez and Zhukov, 1996; Zhukov et al. , 2000b). The softest magnetic
properties, such as magnetic permeability, could be observed in the nearly-
zero magnetostrictive alloys (Zhukov et al. , 2000b). The stresses introduced
by the glass coating give rise to an additional magnetoelastic energy
confribution deteriorating the soft magnetic properties. Consequently, the low
field magnetic properties are expected to be improved by adequate treatment,
resulting in the internal stress relaxation by heat treatment or chemical etching
of the glass coated. For instance, it has been reported quite recently, that
initial permeability can be enhanced significantly by adequate heat treatment
(Zhukov et al. , 2000b). Effect of conventional and magnetic field annealing on
= 0.08, 0.09
hysteresis loop of (Co 1- x Mn x )75 SilO 8 15 (x = O. 09 and O. 10) has been
recently studied (Zhukov et al. , 2000a; Zhukov et al. , 2000b). Most relevant
changes were observed in the samples with x = O. 08 submitted to a set of
different heat treatments (conventional CA and magnetic field FA treatments)
up to 200 C."C. Figure 5. 13a shows as an example of the hysteresis loops
obtained for the sample x = = O. 08 after TA treatments at T ann == 100, 150 and
200 C.
"C. Consequently changes of such magnetic properties as coercivity
(He)' remanent magnetization (J.ioM
(He), (110M,),r ), initial magnetic permeability (1115)
(J.i15)
and magnetic anisotropy field (H k ) with annealing temperature are shown in
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 143

Figs. 5. 13c, d and e. As can be seen, a slight increase of H k is detected with


annealing temperature, while the values of He' 1115 and 110 M, were almost
unchanged with the heat treatments. The axial field annealing FA at the same
annealing parameters produces much more significant changes in the hysteresis
loops (see in Fig. 5. 13b). As can be observed in this figure, field annealing
results in significant changing of the hysteresis loops. These changes become
most significant at T ann ann above 100 C. An increase of the magnetic
permeability, 1115' coercivity, He' remanent magnetization 110 M, and a
decrease of the anisotropy field H k (Figs. 5. 13c, d and e) with T annann have been
observed. It is noticeable to remark that the magnetic permeability, 1115'
strongly increases at T Tann~150C
ann ~ 150 C (Fig.5.13c).
(Fig. 5. 13c) .
These changes observed after field anneal
annealinging should be connected to the
induction of magnetic anisotropy with easy axis along the longitudinal direction
of the microwire. The values of the anisotropy field, H k , after FA as a function
of the annealing temperature (annealing time as parameter) are represented in
Fig. 5.13d. t:..H k has been calculated as the
5. 13d. The field of induced anisotropy, t:.H

~l::zr=:1
-300
~:L2E::1
1:zt=1 =;[:2=:1
-150 0 150 300 -300 -150 0 150 300

p~l:;F=] )[:;t:
)J:::2E:1 p~[:zt:J ,1

~:[~~1~ jj 150 ~
~

l~~;[050:150 ISO ~
-300 -150 0 150 300

1~:L:ZE:J
~~[=zE=~
-300 -150 0 150 300 -300 -150 0 150 300

~::r 200'C )jl,


~:L:ZE:J
~}t:d:::J
-300 -150 0
H(A/m)
150 300
~l200~,
~:~ -, ,)j?
-300
IT
-150 o0
H(A/m)
I

150
I

300
1

(a) (b)

Figure 5. 13 Effect of conventional (CA) (a) and magnetic field annealing (FA) (b) on
hysteresis loops of (COO 92 MnO.08 )75 SilO
(COO.92 '5 microwire at different annealing temperatures,
Si,o 8B15
dependencies of initial magnetic permeability, 1-115' 1-1,5' (c) anisotropy field, H kk ,, (d) and
1-10 M" and coercivity, He' (e)
remanence, 1-10
144 Arcady Zhukov and Julian Gonzalez

FA: ~
-0-0.5
-o-O.5hh
12000 -+-1.0 h
-+-1.5 h

~ 8000
:::
:::l..
:::l

4000

0 50 100
TannCC )
(c)
250

200
CA ...A
__ 0.5h
__ 0.5h
$~ 150
E ISO .......
--1.Oh 1.Oh
FA ......
-+-1.5 1.5 h
~
=e 100 ...... -0.5h
~ -1.0
""'l- -1.0hh
-0- -1.5
-1.5hh
50

0 50 100 150 200


Tann (OC ))
TannCC
(d)
0.5 He: __ CA 40
...... FA 35
0.4
lann=0.5
t ann =0.5 h
30
E:
f=" 0.3 25
25.F :
~
~
::
::: 0.2
J1IJoM -0- CA
oM,:r : -0- $
u
u
-o-FA
-O-FA ::r::
20 :t:

0.1 15

10
0 50 100 150 200
Tann ('C )
TannCC
(e)

Figure 5. 13 Continued from page 143.

difference between the anisotropy field before and after treatments. In the
case of FA treatment t:.H k is of around 120 Aim.
It must be noted that the moderate value of the annealing temperature
cannot assume a pair ordering mechanism for the explanation of the origin of
such anisotropy. It seems reasonable to consider that this anisotropy is
developed by the combined effect of the high internal stress and the magnetic
field applied during the treatment. These results are consistent with those of
stress + longitudinal field induced anisotropy in amorphous ribbons of similar
= 200'C, aann == 500 MPa and H L == 4000 Aim) (Gonzalez,
composition (T ann =
1990c; Zhukov et al. , 2000a; Garcia et al. , 2000). The values of induced
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 145

anisotropy in amorphous microwires are, therefore, quite similar to those


obtained in the ribbons with similar annealing parameters. In addition, this
kind of annealing can develop magnetic bistable behavior in the microwire (for
annealing time), which could be associated to the formation of the "inner
axially magnetized core" region with longitudinal anisotropy. As a
consequence, the coercive field and remanence increase after annealing,
anneal ing,
which could be interpreted by consider the the propagation wall mechanism for
the aforementioned core. Therefore, when the longitudinal induced anisotropy
increases, the magnetic field necessary to move the wall from one end to the
other (which is essentially the coercive field) also increases.
We consider, that the application of the longitudinal magnetic field during
annealing favors the formation of the internal axially magnetized core with
longitudinal magnetic anisotropy. In the mean time, both conventional
annealing as well as chemical etching treatment result only in the internal
stress relaxation. It reflexes particularly the tensor character of frozen-in
internal stresses with important contribution of all components: axial, radial
and tangential.

5.4 Magnetostriction

5.4.1 Amorphous Materials


As has been mentioned, magnetic anisotropy is macroscopically negligible, its
stress derivative defines the magnetostriction coefficient, which depends on
alloy composition. The saturation magnetostriction constant of Fe-base alloys
exhibits large and positive values, while Co-rich alloys have negative
magnetostriction. Vanishing magnetostriction is found for those alloys
containing a ratio of Fe to Co content of about 5 %. The microscopic origin of
magnetostriction is the same as that of magnetic anisotropy and it can be
understood by studying its temperature dependence. From such a study.
study, it can
be determined that single-ion anisotropy is the responsible mechanism for the
magnetostriction of Fe-rich alloys. Deviations from this mechanism are found in
Co-rich alloys in which a significant component corresponding to a two-ion
mechanism is presumed. In fact.
fact, nearly-zero magnetostriction Co-rich alloys
can exhibit a change of sign of magnetostriction well below Curie temperature
(0' Handley and SullSullivan. al . 1984; Lachowicz and
ivan, 1981; Hernando et al.,
Siemko.
Siemko, 1989).

5.4.1. 1 Thermal Dependence of Saturation Magnetostriction


The magnetostriction constant depends first on composition such as has been
146 Arcady Zhukov and Julian Gonzalez

40

x
~ 20
'-;;,
~

o 200 400
T(K)
(a)

5

5:4
~
~ 3

o 200 400 600


T(K)
(b)


-2.5

00
"0 -5 0

x
~
~

-7.5

-10

o 100 200 300


T(K)
(e)
(c)

5. 14 Thermal dependence of the saturation magnetostriction, A"


Figure 5.14 A" of (a)
Fen5 Cu, Nb3 Si'35 8 9 (Fe-rich) positive magnetostriction; (b) C07l1
Si'3.5 B 71.' Feu Si Si'2 8'2
,2 B '2
nearly-zero magnetostriction and (c) C0 75 Si'5 B,o
Si,5 8 10 (0) and CO'5 Ni 30 Si'5 B,o ( )
Si ,5 8'0
negative magnetostriction samples.
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 147

mentioned. The magnetostriction constant also depends on temperature,


although an experimental study of such a dependence is strongly affected by
structural relaxation contribution in the case of nearly-zero magnetostrictive
compounds (Co-rich containing small amount of Fe) (Hernando et al., 1983;
Barandiarcm et ai.,
Barandiaran aI., 1987; Gonzalez and De Lacheisserie, 1989). Figure
5. 14a-c
14a - c presents the temperature dependence of a Fe-rich (a), CoFeSiB (b)
and Co-rich (c) alloys exhibiting positive, nearly-zero and negative
magnetostriction, As' (at room temperature) respectively.
These thermal dependencies of A Ass have been successfully analyzed in
terms of single-ion and two-ion contributions according to the expression
(Callen and Callen, 1965; De Lacheisserie and Krisnhan, 1984):

(5. 10)

where 1 5 / 2 is modified hyperbolic Bessel function of the inverse Langevin


/5/2

function, L -) -1 (m) (0' Handley, 1978), and m is the reduced magnetization

MCT)/M(O). The function 15/2~m3


/5/2~m3 at lower temperatures and ~(3/5) m 2
near Curie temperature.
y=A.!1
For checking Eq. (5.10) we can plot Y=A s //5/2[L
512 [L -l(m)]
-1(m)] against x=
m 22///5/2[L
15 / 2 [L -)
-1 (m)] where x will vary from unity at 0 K up to 5/3 at Curie

temperature (Callen and Callen, 1965). As a consequence, a linear


dependence can be assumed, namely:
(5.11)
(5.11>
where A1 and A2 A 2 are, respectively, the one-ion and two-ion contributions to
the magnetostriction AsA S at 0 K.
The data taken for Fig. 5. 14a for the Fe-rich alloy (Murillo et al., 1993)
are analyzed in Fig. 5. 15 in terms of y
5.15 (x). A single-ion contribution is inferred
y(x).
from the horizontal part which corresponds to the range of low temperature
<
with xx<1.3,
1. 3, i.e., <
i. e., T<350
T 350 K. However for x> x>1.31. 3 a strong deviation from
the single-ion behavior can be observed. Therefore, the single-ion model

45

~
x
<~ 35
3S
~

30 1.0 1.1
1.1 1.2 1.3 1.4 1.5
r=m 2/J512
T=m S12

Figure 5. 15 Analysis of the temperature dependence of the saturation magnetostriction


according to Eq. (5. 11) in the positive magnetostrictive amorphous sample.
(5.11)
148 Arcady Zhukov and Julian Gonzalez

proposed for iron-rich amorphous alloys seems to be valid at low temperatures


in agreement with earlier studies that reported on these kinds of alloys. The
disagreement of this model above 350 K could be related to a two-ion
contribution.
According the experimental data of As ( T) in the CoFeSiB amorphous
alloy (Gonzalez and De Lacheisserie, 1989), the dependence of y is
parabolic (see Fig.5. 16). Within the experimental uncertainties, dy/dx is
Fig. 5.16).
null at 0 K. This means that below 200 K the single-ion contribution dominates
the temperature dependence of As. As' Above this temperature, the curvature can
not arise from a two-ion coupling which cannot be deduced below 200 K. This
curvature could be assigned to a temperature dependence of A 1 as a result of a
competition between the negative one-ion contribution of Co-atoms (A?O (A?o =
- 4x
-4 X 10- 6 ) and the positive one-ion contribution of Fe-atoms (II ~e =
(A~e
+ 4 1 x 10-
+41 6
10 - 6).
). When the sample is heated, some reversible structural changes

occur inducing a very small positive change for IIi\ ~e and/or i\?O,
A?O, thus giving a
much more important relative change for:


25

20

II

'"

1.50 5/3

Figure 5.16 Analysis of the temperature dependence of the saturation


magnetostriction according to Eq. (5. 11) in the nearly-zero magnetostrictive
amorphous sample.

(5. 12)
(5.12)

Therefore, these zero magnetostriction alloys are in fact a very sensitive


probe for any slight
sl ight structural changes.
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 149

Data of As ((T)
n for Co-rich negative magnetostriction of Fig. 5.17 (De
Lacheisserie and Gonzalez 1989) are analyzed also according to Eq. (5. 10). 10) .
For the sample CoSiB such data are perfectly consistent at a low temperature
range reported in (Herzer, 1990a) with those at high temperature data (T>
= 0, they prove that the single-
300 K) from Vazquez et al. (1986). Since 71A22 =
ion model is relevant for describing the magnetostriction of a pure Co-metalloid
amorphous alloy, as was the case of Fe-metalloid alloy, discussed above. It
must be mentioned that the idea of a noticeable two-ion contribution to II Ass in
Co-metalloid alloys was arising from some papers (Jagiel
(Jagielinski
inski et al., 1977;
O'Handley, 1978; Murillo et ai.,
al. , 1993).

1.2 1.4 1.65/3


:I t/,
T=O:
I
i:
/ I
T=7~
I 7< I I
-2.5 : -h +~ I

f!t.++ -tI' 0\-


l-tt--t..J
:I +1t-+++
I
I
-5.0
~oooo 0 00

o
-7.5
~
0

-10.0

Figure 5. 17 Analysis of the temperature dependence of the saturation magnetostriction
(5. 11) in the negative magnetostrictive amorphous samples.
according to Eq. (5.11)

5.4. 1. 2 Stress Dependence of Magnetostriction in Nearly-zero Magnetostrictive


Alloys
Since 1986 it was well established that saturation magnetostriction in metallic
glasses depends on the applied tensile stress, a. This phenomenon was
observed in very low magnetostrictive metallic
metall ic glasses. This dependence,
As ( a), was experimentally found by different authors (Szymczak, 1987;
Herzer, 1990b; Hernando et al. , 1990; Fahnle
Fiihnle et al. , 1991; Hernando et al. ,
1992) to be of the form:
As(a) = 71(0)
A (0) + Aa (5.13)

A (0) is the saturation magnetostriction constant when the applied stress


where II
a is zero, and A is the negative coefficient which ranges from - 6 x 10- 10 10 to

-1 X 10- 10 MPa- 1 .
Figures 5. 18a and b present results given by Blanco et al. (1993), as an
150 Arcady Zhukov and Julian Gonzalez

example of the As A 5 ((J)


( a) of nearly-zero magnetostrictive amorphous
alloys, on the variations of A (0) and A coefficient, respectively, of the
(Co09sFeo.os )72sSi12.SB1S amorphous wire (thermally treated by current
(Co095Feo05)725Sil25B,5
anneal ing) as a function of the current density. Regarding the evolution of
A 5 (0) (Fig. 5. 18a) results are similar to those reported on thermal
As
dependence of nearly-zero magnetostriction amorphous alloys (Gonzalez and
De Lacheisserie, 1989). This As A5 (0) behavior could be accounted by
considering the coexistence of two magnetostrictive phases of different signs
and strengths (Barandiaran et al. , 1987; Lachowicz and Siemko, 1989).
4.5

4.0

3.5
~
'0
l'
I
::
0
3.0
~
~"
~

2.5 31.0 A/mm 2


31.0A/mm
o 34.5 A/mm 2
2.0 38.5 A/mm 2
o 40.8A/mm 2
1.5
0 10 20 30 40 50
tf ann (min)
(a)

31.0 Ahnm 2
31.0Ahnm
o 34.5 A/mm 2
-1.0
A/mm 2
38.5 Ahnm
40.8 A/mm 2
o 40.8A/mm
_I~ -1.2
o
"0
S -1.4

~x
~
":

-1.8 L -_ _L
l --_
_------
--- ,,L
l -_
-_ _ _L:L-_
-_ _l --
_L _ --
_ --
-=- J'
o 10 20 30 40 50
tf ann (min)
(b)

5. 18 Evolutions of the coefficients i\ s (0) (a) and A (b) as a function of the


Figure S.
(annealing parameters of the COO 94 Fe006 )72.5
COO. 94 8'5 Si
)72.58,5 ,2 . 5 amorphous wire treated by current
Si,2.5
annealing.

Different microscopic mechanisms have been invoked to explain the origin


of the parameter A. Fahnle et al. (1991) correlate A with the second strain
derivative of the local anisotropy coefficients which are determined by the
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 151

local symmetry and chemistry. Therefore, the thermal treatments giving rise
to irreversible phase transformations in the local symmetry of the amorphous
structure are expected to affect the A value. It is to be noted that different
experimental hints of such types of transformations have been observed for
Co-rich metallic glasses (Carb
(Corb et al. , 1983).
The model developed by Szymczak (1987) describes the stress
dependence As A s as a consequence of the bond orientational anisotropy induced
by the stress. Therefore, in the framework of Szymczak's model, thermally
activated processes must be invoked in the mechanism giving rise to the stress
dependence of As. Moreover, it is expected that the action of the tensile
stress should be drastically affected by the strength and orientation of any bond
anisotropy induced previously at higher temperatures.
Hernando et al. (1990 and 1992) have analyzed the As ( a) behavior in
amorphous ribbons taking into account the fluctuations of the local anisotropy
and, therefore, the local magnetostriction. When the local magnetostriction
fluctuates with a correlation length larger than the exchange correlation length,
As varies with the applied stress. Moreover, in this last model, the influence
of the thermal treatments on the local magnetostriction fluctuations should be
reflected in a similar influence on the coefficient A.

5.4.2 Nanocrystalline Materials

A~ff, observed in nanocrystalline alloys at


The effective magnetostriction, A;",
different stages of crystallization has been interpreted as a volumetrically
weighted balance among two contributions with opposite signs originating from
(A ~r) and residual amorphous matrix (A
the bcc-FeSi grains Ci\ Ci\ ~)
~m) according to
(Herzer, 1993):
(5. 14)

where p is the volumetric fraction of the crystall ine phase. Therefore,


assuming negative and positive signs for the nanocrystalline and amorphous
phase respectively, the variations of A A;"~ff (including the change of sign
observed in some nanocrystalline compositions) can be interpreted as a
consequence of the variations of p parameter. Although this simple
approximation gives the qualitative explanation of the effective
magnetostriction in Fe-based nanocrystalline alloys (Herzer, 1993),
nevertheless, it does not consider many effects occurring in the real materials.
More exact calculations take into account that the magnetostriction of the
residual amorphous phase is not constant but depends on the crystalline
fraction due to the enrichment with Band Nb (Muller and Mattern, 1994;
Twarowski et al. , 1995). Consequently, Eq. (5. 14) can be modified in the
CTwarowskii et al. , 1995):
form (Twarowsk
(5. 15)
152 Arcady Zhukov and Julian Gonzalez

where k is a parameter which expresses changes of the magnetostriction in the


residual amorphous phase with evolution of the crystallization. In many cases
even this model does not fit the experimental results demonstrating that the
effective magnetostriction in nanocrystalline material cannot be described by a
simple superposition of the crystalline
crystall ine and amorphous components (Muller and
Mattern, 1994). In the case of FeZrBCu
FeZr8Cu nanocrystalline system in which the
bcc-Fe phase is formed, the model described does not fit the experimental
data even through the A ~m ( p) dependence as was shown by Slawska-
Waniewska and Zuberek in (1996a, 1996b). They considered an additional
contribution to the effective magnetostriction, which arises from the enhanced
surface to volume ratio describing interfacial effects (Slawska-Waniewska
et al., 1996a and 1996b). Therefore, Eq. (5. 15) of the effective
magnetostriction could be approximated by:
A~ff
A;" = pA~r
= pA;r + (1 - p)(A~m + kp) + pA~S/V (5.16)
where the last term describes the effects at the interfaces and depends on the
surface to volume ratio S / V for the individual grain, as well as on the
magnetostriction constant A~ which characterizes the crystal-amorphous
interface.
Equation (5.16) is the basic dependence which can be used to interpret
the experimental data on the effective magnetostriction in Fe-based
nanocrystalline alloys at different stages of crystallization. The composition of
the Fe (Si)
( Si) grains (depending on the annealing temperature) should be
considered giving rise to different values of the magnetostriction constant for
the crystalline phase. The appropriate values of A;rA~r can be obtained from the
compositional dependence of the saturation magnetostriction in polycrystalline
ex-FelOo.-xSix shown in Fig.5.19 (Yamamoto, 1980; Herzer, 1993). Thus, the
cx-FelOo-xSix
first term in Eq. (5. 16) can be treated as the well defined one.
(5.16)

'f'
o
~
x
<-<.~ -10

o 5 10 15 20
Si in a-FeSi.(at.%)
u-FeSi.(at.%)

Figure 5. 19 Saturation magnetostriction of the polycrystalline a-Fe100-


a-Fe100-'x Si,
Six..

Figure 5.20 presents the crystallization behavior and accompanying


changes in the magnetostriction of classical Finemet (Fe73.5 CUT Nb 3 Si 135
(Fen5 CUI 13 . 5 B
89 )
alloy published by Gutierrez et al. (1995). The analysis of these data
according to Eq. (5. 16) allows CD estimation of the contribution from the
crystall ine phase (see Fig. 5. 8a where the values of A;r
crystalline A~r were found from the
combined Figs. 5.19
Figs.5. 19 and 5.20a, CV fitting of the experimental (i\;"
5. 20a, and ~ CA~ff - pA~r)
pA;r) on
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 153

p dependence to Eq. (5. 16). The results, both experimental points and the
fitted curve (solid line) are shown in Fig.5.20b.
~18.0
~
~
18.0
?f,
~ 17.5
~
'{j 17.0
u...
lJ..
6
c:
c::
.5
16.5
16
u; 16.0
Vi
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Crystalline fraction
(a)

~ 20
i'
'f
::S 15
x
2S
~

~V>
1E", 10
~
5
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Crystalline
Crystall ine fraction
(b)

Figure 5.20 Si content in (X-Fe(Si)


o:-Fe(Si) grains (a) and magnetostriction of the
Fe73sCu,Nb3Si,3sB9
Fe735Cu,Nb3Si135B9 alloy (b) versus crystalline fraction.

Assuming spherical ex-Fe(Sj)


o:-Fe(Si) grains, with radius R, the last term of Eq.
(5.16) can be expressed as 3A~/R, 311~/R, and the magnetostriction constant which
describes the interface A~ 11~ can be estimated. For the soft nanocrystalline alloys
(Finemet and FeZrBCu alloys) (MUlier (Muller and Mattern 1994; Herzer 1993;
Slawska-Waniewska et aI., 1996a) R = 5 nm and thus 11~ A~ has been found to
vary in the range 4.5 X 10- 6 -7. -7.11 x 10- 6 nm. These values result to be one
order of magnitude smaller than values of the surface magnetostriction obtained
in multilayer systems. However, investigations of Fe/C multi layers have
shown that, not only the value, but also the sign of the surface
magnetostriction constant depends on the structure of the iron layer and it has
been found that for crystalline iron A~(bcc-Fe)
1I~(bcc-Fe) =45.7
=45. 7 x 10- 6 nm, whereas for
the amorphous iron A~ 11 ~ (am-Fe) = - 31 X x 10- 6
10 -6 nm (Zuberek et aI., 1993).
Thus, the value of the interface magnetostriction obtained in the
nanocrystalline
nanocrystall ine systems is within the range of the surface magnetostriction
constant estimated for thin iron layers. It should be noted that, contrary to Fe/
C multilayers,
multi layers , in the nanocrystalline materials, both the crystalline and
amorphous phases are magnetic and they are coupled through exchange and
dipolar interactions. It must thus be expected that the magnetic interactions in
the system can affect the magnetoelastic coupling constant at the grain-matrix
interfaces. In addition, the structure and properties of the particular surfaces
being in contact as well as local strains at the grain boundaries should also be
154 Arcady Zhukov and Julian Gonzalez

cons idered.
The problem of the surface/interface magnetostriction requires however
further studies which, in particular, should include measurements at
temperatures above Curie point of the amorphous matrix where only
ferromagnetic grains should contribute to the effective magnetostriction,
simplifying a separation between bulk and surface contributions.

5.5 Giant Magneto Impedance Effect

The study of the giant magneto impedance (GMI) effect became a topic of
intensive research in the field of applied magnetism during the last few years
(Makhotkin et al.,
aI., 1991; Pan ina and Mohri, 1994; Beach and Berkowicz,
1994; Machado et al. , 1995; Ciureanu et al. , 1996). This GMI effect consists
of a large change of the electric impedance of a magnetic conductor when it is
subjected to an axial dc magnetic field. It has been recognized that the large
sensitivity of the total impedance of a soft magnetic conductor at low magnetic
fields and high frequencies of the driven ac ae current originates from the
dependence of the transverse magnetic permeability upon the de dc magnetic
field and skin effect. The main interest of the GMI effect is related to the high
sensitivity of the impedance to an appl ied magnetic field, achieving up to
applied
300 % relative change of impedance in conventional amorphous wires with
vanishing magnetostriction (see Fig. 5. 21 for the (COO 94 Fe006 )725 )72.5 B 15 Si 125
12 . 5
conventional amorphous wire). In the meantime, a generally inferior GMI
effect has been observed for the other amorphous magnetic materials, such as
amorphous ribbons and microwires, where the GMI ratio is generally less than
60 % (Sinnecker et al., 1998). 1998) . Circular domain structure with high
circumferential permeability proved to be very favorable for highest GMI effect
(Pan ina and Mohri, 1994; Beach and Berkowicz, 1994). Such domain
configuration is typical for the nearly-zero magnetostrictive amorphous wires of
Unitika LTD (Humphrey et ai., al., 1987a; Pan ina and Mohri, 1994; Atkinson and
Squire, 1997). On the other hand, these amorphous wires with vanishing
magnetostriction constant present the best magnetic softness (Humphrey
et al. , 1987a). It has been clearly demonstrated by Mohri (Yoshinaga et al. ,
1999) by chemical etching of 30 IJm (COO 94 Fe006 )725 )72.5 Si 125 B 15 wire in 10%

HN0 3 , that the removing of the external layer with high transverse
permeability results in simultaneous degradation of its magnetic softness,
(tl.Ew/E
decrease of MI effect (t:::.E w)) and increase of the squareness ratio and
w / Ew

coercivity. In addition, it was recently demonstrated that the appl applied


ied tensile
stresses can significantly modify the magnetoimpedance response of
conventional amorphous wires and glass coated microwires (Humphrey et al. ,
1987a; Atkinson and Squire, 1997; Yoshinaga et ai., aI., 1999; Aragoneses
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 155

et al. , 2000).
300
_ 1=5rnA
1=5mA
250
-+-
...... 1=\ rnA
1=lmA
~
~ 200
tN~ 150
~ 100
50
a0
a0 500 1000 1500
\500
H(A/m)
H(A/rn)

Figure 5.21 Axial field dependence of the GMI ratio in as-cast (COO
CCoo9,Feoo6)725B'5Si'25
94 FeO.06 )725 8 15 Si 125

amorphous wire.

As has been mentioned, initially the GMI effect was interpreted in terms
of the classical skin effect in a magnetic conductor assuming scalar character
for the magnetic permeability, as a consequence of the change in the
penetration depth of the ac current caused by the dcde applied magnetic field.
The electrical impedance, Z, of a magnetic conductor in this case is given by
(Panina and Mohri, 1994; Beach and Berkowicz,
Berkowicl, 1994):

= R dc krJ o(kr)/2J,(kr)
Z = (5. 17)
(5.17)

= ( 1 + j) / 5 where J o and J, are the Bessel functions, r wire's radius


with kk=(1+j)/5
and 5 the penetration depth given by:
(5.18)
(5. 18)
where a is the electrical conductivity, f the frequency of the current along the
sample, and J.I~
J.1 the circular magnetic permeability assumed to be scalar. The
de applied magnetic field introduces significant changes in the circular
dc
permeabil ity, J.IJ.1 ~.
. Therefore, the penetration depth also changes through and

finally results in a change of Z (Pan ina and Mohri, 1994; Beach and
(Panina
Berkowicz, 1994). Recently this"
Berkowicl, scalar" model was' significantly modified
this "scalar"
taking into account the tensor origin of the magnetic permeability and magneto-
impedance (Aragoneses et al. , 2000; Makhnovsk iy et al. , 2001). Generally,
the magneto-impedance characteristics can be obtained from two separate
voltage responses with changing external applied field Hex: CD voltage V Vz
across the MI sample and (2) ~ voltage V c in the external coil
Vein coi I (mounted around
the MI sample). These two voltages can be expressed in terms of components
A

of the impedance tensor s' s'(Zhukova


(Zhukova et al. , 2002).
The induced voltages are determined by the (tangential) surface value of
the variable electric field e" which is related to the surface magnetic field h,
via the surface impedance tensor s' (Makhnovskiy et al. , 2001; Zhukova et
A

al. , 2002):
156 Arcady Zhukov and Julian Gonzalez

(5.
(5.19)
19)

where the cyl indrical coordinate system (n r' n ~, n z) is used with the vector
cylindrical
n zz directed along the wire axis.
The voltage V z is determined by the surface value of the axial electric
field e z :

Vz =
V e zl = (- s'5'" z~h z + s'5'" zzh ~~)) 1.
I. (5.20)

Here 1 I is the wire length, h ~ (0) = 2 i w / (co)


(a) = (ea) is the field induced by ac current
iw=iowexp(-jwt) (w = 2TIf, f is the frequency), c
i w = i ow exp ( - jwt) in the wire (W=2TTf, e is the
velocity of light, and h zz is the ac axial magnetic field (from the external
primary coil, for example) .
The voltage Vein the external secondary coil (signal coil) can be
obtained in the form (Makhnovskiy et al. ,2001; , 2001; Zhukova et al. ,2002):
, 2002):

jwN
V c -_ - -
-h
2
h z TT(02
ec
2 2
+ 2TToN
TI ( a 2 - 0a ))+ TI aN 22 ( -- s' ~h zz
5'"<pq>h + 5'"~zh~
s' )
~zh ~ . (5.21)

Here the primary coil is wound over the secondary coil, a 0 is the wire radius,
a2 is the secondary coil radius, N 2 is the number of turns of the secondary
02
coil. The first term in Eq. (5. 21) corresponds to the contribution from the
magnetic flux through the air gap between the wire and the secondary coil.
The second term represents the contribution from the magnetic flux through the
wire expressed via the tensor components.
It must be mentioned that each component of the impedance tensor has its
own field characteristic and affects the sensitivity parameters specifically. In
the case of the wire with the helical anisotropy the dc bias field induced by the
current 1 I results in asymmetry in the static magnetic configuration. The
modifications in the impedance plots due to 1 I are of the same kind as those for
the de magnetization loops: with increasing 1 I the hysteresis loop becomes
narrower and asymmetrical. This shows clearly that the asymmetry in the
impedance-field (MI) behavior is related to the static magnetic structure that
originates from a combination of the helical anisotropy and dc circumferential
field (Makhnovskiy et al. , 2001; Zhukova et al. , 2002) (see all components
of the GMI tensor in Fig. 5.22)
5. 22)..
On the other hand, it was demonstrated (Blanco et al., 1999a and
1999b; Tejedor et al., 1999; Blanco et al., 2000) that the application of
tensile and torsion stresses results in significant changes of the GMI effect
through the modification of the domain structure under such stress. Therefore,
the stress impedance has been introduced in order to characterize the change
of the electrical impedance under the stress (Gonzalez et al., 1994; Shen
et al. , 1997; Blanco et al. , 2000; Blanco et al. , 2001). Particularly, it was
found that the application of only the torsion without any magnetic field gives
rise to a significant change of the electrical impedance (Shen et al.,ai., 1997;
Blanco et al. , 2000; Blanco et al. , 2001). When the torsion dependence of
the impedance was investigated the torsion impedance ratio (TI), (6,Z / Z) ~~,,
CTI), (I:lZ/
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 157

2.0

1.5

~ 1.0
>"...P
<f 0.5

o
--40 -20 0 20 40
Hex (Oe)

1.0 ?-~

0.8 ll~o'.\ 1=70mA

J: 0.6 1r ~ \
~
~ 0.4
'-J>
<f 0.2 I 11 \\
\ ~
I;
o ~,.,~.-.... "'~~ ~

--40 -20 0 20 40
Hex (Oe)
2.5

2.0
\ 1=150mA

\
\ \
\
~--F.~oYD
..po'" '-
iJ,.o-O
000...._
~-~
o .;J . . . . . . .- .",....
--4':::-o----2::-:o,----~0----;!20::----4:-:::-0
--4'::-o----2::'-:o~---:':0----:!20::----4'""0
Hex (Oe)

Figure 5.22 >' zz ,, >'


)' II )' ,' >'.z
)'.z components of the GMI tensor measured in
(CO
CC0094
.
O 94 Fe006 )
)725
725 8 15 Si 12 . 5 torsion annealed amorphous wire at bias
125

current of 70 mAo

has been determined by the expression:


(/).Z/Z)s =
(/::,.Z/Z)~ = [Z(~appl)
[Z(~apPI) - Z(~max)J/Z(~max) (5.22)
where ~max is the maximum torsion stress.
stress, ~apPI applied for the determination of
this (/).Z ~. Measurements of TI
(/::,.Z / Z) s. Tl with and without a longitudinal magnetic
field, H, can be made.
field.
Generally,
Generally. torsion stress dependence of total impedance.impedance, Z. Z, and
(/::,.Z / Z) s~ have a shape of decay with applied torsion.
(/).Z torsion, ~.
~, beginning from some
position corresponding to their maximum. This decay has saturation indicating
158 Arcady Zhukov and Julian Gonzalez

that after appl


application
ication of large enough torsion Z and (/1Z (/::,.Z I Z)
Z ) s{ are almost
independent on ~. The position of the maximum can change after different
treatments. It is worth mentioning that as-cast (Coo.94Feo 06)725B15Si,25
amorphous wire has an asymmetric Torsion Giant Impedance (TGI) effect and
such asymmetry can be substantially eliminated by current annealing or
substantially enhanced by the current annealing under torsion (Fig. 5. 23a). 23a) .
The asymmetrical character with respect to zero torsion found in the
as-quenched state is more significant in the torsion annealed sample with
rather sharp maximum of (/1Z (/::,.Z I Z) sm
~m around 200 % at ~ = 20n 20TT rad/m. By
applying a magnetic field (15 Aim), the maximum change on the /1Z /::,.Z I Z s~ ratio
results to be around 225 % for the same ~~appi appl = 20n
20TT radl
rad/m.m. Both curves are
also plotted in the Fig.
Fig.5.23a.
5. 23a.
300

250 x x

200 x
~

'~
<J?
~ 150
~
N
t!
N
3~ 100
50

0
-30 -20 -10 o 10 20 30 40
~ (mad/m)
(]trad/m)
(a)

300
e
250 -e- Without pre-annealing
- 0 - After pre-annealing
~

e
0~
~

E 200
"'
>V'
0
~
~
S 150

100 e

0 20 40 60 80 100 120
lann (min)
(b)

(/!,.Z / Z) {~ ratio of (COO 9'


Figure 5.23 The (1:lZ/ 94 Fe006
FeO.06 )725 I5 Si
B15
)72.5 8 ,25 amorphous wire in as-cast
Si,25
state (.), pre-annealed (0) and pre-annealed and torsion annealed without magnetic field
(0) and under magnetic field of 15 A/m( X).

A good symmetry was found with respect to ~ =0 for the pre-annealed


Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 159

sample which should be mainly ascribed to the relaxation process connected to


the shear strain-relief in the metallic nucleus. The maximum TI ratio,
(f;.Z / Z) ~m' results to be around 180 % and this (b.Z
(b.Z Z ) ~ ( ~) dependence
(f;.Z / Z)
became narrower when it was compared with the curve of the as-quenched
sample. Such as is observed in the case of the sample treated only under
(Fig. 5. 23a), (f;.Z
torsion annealing (Fig.5.23a), / Z) ~ ( ~) curve of the pre-annealed +
(b.Z/Z)~(~)
torsion annealed sample presents a well defined maximum (similar to the pre-
annealing case) but this peak takes place at ~ = 1 =
hr rad/m, such as can be
ll1T
hel ical induced anisotropy. In Fig. 5. 23b the evolution
expected owing to the helical
of (f;.Z/Z)~m
(b.Z/Z)~m is presented with the annealing time, t ann , measured for the
torsion annealed and pre-annealed + torsion annealed. The largest change of
(b.Z / Z) ~m ratio is observed in the torsion annealed sample to be around 270 %
(f;.Z
without applied magnetic field.
After Joule heating under applied torsion, (b.Z (f;.Z / Z) ~ has a tendency to
achieve finally a sharp and rather asymmetric shape with a maximum at certain
torsion, ~m'
~ m' In the case of the sample annealed under torsion, the position of
the maximum, ~m' of (f;.Z/Z)~ratio
(b.Z/Z)~ratio depends on time of current annealing
(Fig. 5. 24) as well as on torsion applied during this annealing at given
annealing time (Fig. 5. 25).
300 as-cast
o l. nn =1 min
tann=1
250
.. tlann=5
ann=5 min
v tlann=
ann =15min
15 min
200 tann =105min
lann= I 05 min
~
;?
d'(
~
)0 150
t::!
~
~ 100

50

-20 -10 o 10 20 30
~ (mad/m)

Figure 5. 24 Effect of the current annealing time, t aaoo


"",' on the (/:;Z/ Z)~ (0
(I:J.Z/ Z), () dependence.
Lines are drawn as guides for the eyes.

To explain the TGI effect, it is necessary to consider the fact that the
applied torsion strain induces a helical magnetic anisotropy of magnetoelastic
character. Such helical anisotropy should be in competition with that ascribed
to the complex internal stresses introduced during the fabrication process.
Consequently, the circular magnetic permeability is enhanced by the torsion.
That contribution should be connected with the tensor character of the magnetic
permeability. Consequently, the contribution to the circular permeability due
to the helical anisotropy should be negligible when the maximum (b.Z / Z) ~ is
(f;.Z/Z)~
achieved.
160 Arcady Zhukov and Julian Gonzalez

350 as-cast
D.
b. 81t rad/m
8n
300 T 121t rad/m
12n
~ 250 o 201t rad/m
20n

~uJ'
u.f'
200
281t rad/m
28n
~

~,
~. 150
- 100

5~1..'~~~~~S!rl~~~~
5~1.1II'~ii~~~S!rl~~~~
-20 -10 0 10 20 30 40 50
~ (nrad/m)
(mad/m)

Figure 5. 25 Effect of the torsion applied during torsion annealing on the (t:>.Z / Z), ( P
C/:;.Z/Z),CE,)
dependence. Lines are drawn as guides for the eyes.
dependence,

The mentioned slight


sl ight asymmetry and improvement of t,.Z /::,.Z / Z ~m under low
magnetic field could indicate the fact that spontaneous helical anisotropy due to
the fabrication procedure cannot be compensated only by the torsion owing the
complexity of such internal stresses distribution.
As it was reported by Gonzalez et al. (1994), torsion annealing induces
hel ical anisotropy in the amorphous wire investigated in this work. As twisting
helical
the wire with helical induced anisotropy, a magnetoelastic anisotropy of helical
character is also induced on the wire. For values of ~ corresponding to the
maximum of (/::,.Z Z ) ~ both anisotropies can interact with and enhance the
(t,.Z / Z)
circular magnetic permeability because of non-diagonal terms of the
susceptibility tensor. It is interesting that even in this state the axial magnetic
(t,.Z// Z) ~ ratio confirming strong non-diagonal character of
field still improves (/::,.Z
tensor of magnetic susceptibility.
The pre-annealing results in relaxation of internal stresses induced by the
fabrication procedure. In contrary, the torsion annealing induces helical
anisotropy with a narrower distribution of intensity resulting that the shape of
(/::,.Z / Z) ~ presents a peak with better resolution than compared to that of the
(t,.Z
as-quenched sample (Fig. 5.23) 5. 23)..
Finally, the effect of the pre-annealing on the torsion dependencies of
(t,.Z / Z)~
(/::,.Z/ Z) ~ ratio found in Fig. 5.235. 23 should be ascribed to different contributions
(annelastic and plastic) invoked (Gonzalez et al., 1994) for the hel ical
induced anisotropy in amorphous alloys. It has been assumed that the
annelastic component has a reversible character in the sense that it disappears
after pre-annealing while the plastic component is permanent. This last plastic
component is predominant in the pre-annealed samples at elevated
temperatures as it is the case of the pre-annealing at 550 mA, 2 minutes
carried out in this work. Therefore, the difference on the torsion angle for the
maximum value in the (/::,.Z (t,.Z / Z) ~ curves between the sample treated under
torsion annealing (20lT (201T rad/m) and the pre-annealed + torsion annealing (11lT (1 hT
rad/m) could be related with the annelastic component of the helical induced
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 161

anisotropy. In this way, the helical hel ical anisotropy induced in the pre-annealed +
torsion annealing
anneal ing sample is mainly of plastic character. In addition, the
distribution of this plastic component should be narrower than in the case of the
torsion annealed sample and to show a more significant (I1Z (b,.Z / Z) ~ effect when
the spontaneous hel ical induced anisotropy is compensated by the hel
helical ical
helical
magnetoelastic anisotropy introduced by the applied torsion strain.
Recently GMI effect has been also observed in glass coated amorphous
microwires. The sample composition has been carefully selected among the
C069-xMn6+xSi10815(0<x<1) compositions in order to achieve the
series of C069-xMn6+xSilQBls(0<x<n
best combination of soft magnetic properties (high magnetic permeability, low
magnetic anisotropy field and low coercivity) in the as-prepared state on the
basis of the hysteresis loop measurements.
Regarding Cobeno et a!. al. (1999), an increasing of the magnetic
permeability and coercivity and decreasing of the magnetic anisotropy field
C0 69 - x Mn6+ x SilO 8 15 (0< x < 1) compositions with x at xx<
were observed in C069-xMn6+xSiloBls(0<x<n <
0.75. The hysteresis loop at around x = ~ O. 75 becomes rectangular and the
initial magnetic permeability drastically decreases. Accordingly, the
magnetostriction constant of C0 69 - x Mn6+ x SilO 8 15 samples changes its sign at
C069-xMn6+xSilOBls
around x = = O. 75 (Cobeno et al., 1999), being of negative character at x < <
0.75. Small negative magnetostriction constant can be assumed for the sample
with 6.5 % at Mn. Accordingly, it is expected that the outer domain structure
changes its circular easy magnetization axis (very favorable for the GMI
effect) to the radial one at xx~O. =0.75. 75.
Considering the above mentioned, the sample with x = O. 5 (with total
microwire diameter, 0=16.5 D~16. 5 IJm, and the diameter of the metallic nucleus,
d, of around 7.5 IJm) has been selected for the experimental studies of the
GMI effect. Pieces of 100 mm length have been annealed at 100 C 'c for 1 hour
under axial magnetic field of 14 kA/ kA/m m in order to enhance their initial magnetic
permeability (Zhukov et al., 2000a). According to experimental results
reported by Zhukov et al. (2000a), ( 2000a ), the initial magnetic permeability
drastically increases from 2000 to 12,000 after such magnetic field anneal ing.
The axial field dependencies of the magneto-impedance ratio measured at
different frequencies, f, of the ac driving current amplitude of 1 mA for field
annealed sample are presented in Fig. 5.26. The observed dependencies have
generally similar shape such as was observed in conventional wires (Panina
and Mohri 1994), showing a maximum at certain dc axial magnetic field, H m .
As can be seen from Fig. 5.26, the maximum GMI ratio, (I1Z/Z)m' (b,.Z / Z)m, as well as
Hmm increase with f up to 10 MHz.
The axial field dependencies of the GMI ratio measured at 10 MHz with
different values of the ac driving current amplitude, ampIitude, I, from 0.5 O. 5 up to 2.5 mA
are presented in Fig. 5.27. Non-monotonic dependence of the maximum GMI
ratio, (I1Z
(b,.Z// Z) m' has been observed with a maximum at 1= 1.5 - 2 mA where
Z)m'
(I1Z Z ) m reaches a maximum of around 122%
(b,.Z// Z)m 122 % (at current ampIitude
amplitude of
0.5 mA (I1Z ~ 97 % ). Two different effects should be considered. The
(b,.Z / Z) m =
162 Arcady Zhukov and Julian Gonzalez

150
ISO -e--pl
--e- f=1 MHz
o {=2MHz
-0- f=2 MHz
-...-
-A- /=4 MHz
100 -\7- f=6MHz
-'1- {=6 MHz
-+- p7Ml-lz
-.- j=7MHz
~ ,\..
:.. '\ ++ oo 00 -0- (=10
f=IO MHz
~ V'
o :., \7 +.
++ 000 00
O0 ...A
e \\0 '"
oeO
V'v...+.+..
\7V .+
.
'&A \7\7V'~
...... \7V'
-. bo~"'~ ~
'--
"'-
"'-0
'--0 ______
0"
b AAJt.4l?:;;
~B 0--0_
0-0-
"~<200~=:::::::::~===-+-
.. ----- -+-+-
--ee~- -o=---
... i=iF
=::::::;-i=iF
o
250 500 750
H(A/m)

Figure 5.26 Effect of frequency on l:J.Z/ SilO 8'5


/:;.Z / Z ratio of C0685 Mnu Si,o S15 microwires.

first one is related with the circular magnetization process (i. e., with the
effect of a circular magnetic field, produced by an ac driving current on the
circumferential magnetic permeability). This first effect was previously
observed and described by various authors and has a reversible character
(Vazquez et al., a\., 1998; Aragoneses et al., aI., 2000). Driving current
dependence of the GMI means that we deal with the non-I inear GMI effect,
i. e. , that the domain structure of the outer shell plays an important role. In
H m (observed in Fig. 5.27)
this case a slight increase of Hm(observed 5. 27> can be attributed to
the effect of the circular field amplitude on the circular hysteresis loop. It is
expected that, like in the case of the axial hysteresis loop, the increase of the
magnetic field amplitude results in an increase of the circular coercivity
(Vazquez et al.,aI., 1998). On the other hand, it should be mentioned that the
driving current ampl itude should be related to the current density. In fact a dc
amplitude
current of 5 mA corresponds to a current density, j, of around 110 AI mm 2 for
studied microwire (calculated as j = IlrrR =
I hrR 2 ).
). Such current density could

produce drastic and irreversible deterioration of its magnetic softness due to


the Joule heating (Zhukova et al. , 2001). For instance, experimental results
obtained by Blanco et al. (1998) for amorphous ribbons and wires without
glass coating showed that a current density of around 40 - 50 A/mm 2 results in
a heating of the sample up to around 350 - 400C. Besides, it was
theoretically shown that the glass coating creates an additional thermal
shielding due to the insulating properties of the glass (Chiriac et al. , 1999). In
the case of ac current with amplitude of 5 mA, the Joule heating is localized in
the surface layer and the skin effect should be taken into account to estimate
exactly the heat distribution.
Axial field dependencies of I1Z1 /),Z I Z measured under applied tensile
stresses, (J,a, ranging between 0 and 132 MPa (0 - 2. 9 g of mechanical
=
loading) and at f = 10 MHz and 1= I = 1 mA are shown in the Figure 5.28. In the
unstressed state there is a maximum in (I1Z (/),ZII Z)m
Z ) m for a value of around
130 Aim of axial applied field. This maximum displaces towards larger
applied (/),ZII Z)
appl ied fields and the value of (I1Z Z)m m has a non-monotonic dependence on
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 163

120

~ 80
~
~ 40

o 200 400 600


H(A/m)

5. 27 Effect of driving current amplitude, I,


Figure S. I. on GMI ratio of C068
6S 5Mn6 5Si,oB'5
5Si,o B'5
amorphous microwires.

(]' with a broad maximum at around 60 MPa. The maximum GMI ratio,
(j

(I1Z// Z)
(/1Z m' reaches 130%
Z)m' 130 % at such tension (see the inset of the Fig. 5.28).
5. 28) .
150
210

120 D~~~!
O~;e~~! IflO . -- __ 120 ~
120 ~
d f
d / !or~~:140
j~~140 =~~Z/Z)m
=h~Z/Z)m ~
N
~
~
~

~
::f?
90
_//0
-J/ Q~'
bn 't 0
~
0 60
60 120
120
100
100

/'o~J/
. / If
ID~'/
0

~
~

N / , 0 cr(MPa)
U (MPa)
~
~ 60
60 ~
<I
/1
o
/I!f 00
I*d -
0/*0
Without stress
u=33MPa
cr=33 MPa
~~
1iIi~~
-~.~
30 I. <:f O~~
0 f::,. u=66
.6. MPa
cr=66MPa ~~
~' u=99 MPa
T cr=99
'"
0 o cr=132MPa
u=132MPa
0 200 400 600
H(A/m)
H(A/rn)

Figure S.
5. 28 Effect of (j a on f,Z
/::,.Z / Z C
( H) dependence measured at 10 MHz (a) Ca) and
dependencies of Cf,Z/Z)m
(/::,.Z/ Z)m and H mmon (j a Cb)
(b) of C06s5Mn65SilOB,5
C068.5Mn6.5SilOB,5 amorphous microwires.

Magnetic field, H m corresponding to the maximum of I1Z/ /1Z/ Z shows a


(]'. This tensile stress dependence is rather similar
roughly linear increase with (j.
to that recently reported for conventional (much thicker) wires by various
authors (Knobel et aI., 1996; Atkinson and Squire, 1997). 1997>. These results
should be considered of special worth as regarding the inherent difficulties
associated with the fact of applying a small load on the microwire placed into
the impedance measurement circuit. The tension produced by such a small
load (up to around 3 g) resulted to be big enough owing to tiny microwire
dimensions. It makes the observed effect very sensitive to small mechanical
load. The origin of this dependence should be related to the change of the
domain structure inside the outer shell with circular orientation. As is well
known, applied
appl ied stress introduces a magnetoelastic anisotropy contribution
164 Arcady Zhukov and Julian Gonzalez

which plays a very important role in the magnetization process of the metallic
nucleus, even in this microwire exhibiting extremely low magnetostriction
7
constant (of around --3x3 x 10-
10 -7)) Wobeno
(Cobeno et al.,
al. , 1999).
It is remarkable to mention that according to Pan ina and Mohri (1994),
Atkinson and Squire (1997) and Blanco et al. (1998), the position of the dc
axial field that corresponds to the maximum GMI ratio, H mm', should be
attributed to the static circular anisotropy field, H k . This argument allows us
to estimate the magnetostriction constant using the dependence H m ( a)
presented in the inset of Fig. 5.28 and the well known expression for the stress
dependence of anisotropy field (Knobel et al. , 1996; Cobeno et al., al. , 1999):
= (IJ
As = M s /3)(dH k /da)
(J..looM,j3)(dH (5.23)
The Hm(a) dependence (see the inset of the Fig.5.28) is roughly linear
with a slope of around O. 7 AI (m MPa), that allows an estimation of the
unstressed value of the saturation magnetostriction constant, As.o s.O We have

A s.o:=:::::: - 2 X 10- 7
found that As.o""" 7
,, which is rather reasonable in comparison with the
recently reported values measured from the stress dependence of initial
magnetic susceptibility (A (As.o :=:::::: - 3 X 10- 7
so """
7
for such composition) (Cobeno
et al. , 1999).

600 /=10 MHz.


MHz, i=0.75
/=0.75 mA

- . - p=0.98
-D
- - p=0.816
c ,.-
-e
- .- - p=0.789

a
o 500 1000 1500 2000
H (Aim)
(A/m)

Figure 5.29 Effect of ratio p = d/D on /:;Z/ Z( H) dependence


of C0 67 05 Fe385 Nil 4Bl155Si14
C067.osFe3.8sNit 4B 1 \.ssSi t4 5
sMo 165 amorphous microwires.
M0 16S

Recently colossal GMI ratio (about 600 %) has been observed in


C067.05Fe3.85Ni14Bl1.55Sit4.5Mo165 glass coated microwires after careful
C06705Fe385Ni14B1155Si145Mo165
selection of their composition and geometry (Zhukov, 2002) (Fig. 5.29). The
significant GMI ratio obtained in the glass coated microwire should be
attributed to an adequate processing (composition, geometry, heat
treatment> of the sample that enhances the magnetic permeability, of the
sample. In fact, most experimental and theoretical results pointed out that the
good magnetic softness is directly related with the GMI effect. For instance, it
was theoretically shown by Usov et al. (1998), that the axial dependence of
the GMI spectra is mainly determined by the type of magnetic anisotropy. It
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 165

was shown particularly, that the circumferential anisotropy leads to the


observation of the maximum of the real component of wire impedance (and
consequently of the GMI ratio) as a function of the external magnetic field. In
contrast, in the case of axial magnetic anisotropy, the maximum value of the
GMI ratio corresponds to zero magnetic field (Usov et al., 1998), i. e. ,
results in monotonic decay of GMI ratio with axial magnetic field.
Consequently, non-diagonal components of the magnetic permeabil ity tensor
permeability
and impedance tensor were introduced in (Aragoneses et al., 2000,
2001; Zhukova et al. , 2002; Zhukov, 2002) in order to
Makhnovskiy et al. ,,2001;
describe such circumferential anisotropy. Therefore, for the samples with
well-defined maximum on the axial field dependence of GMI ratio, the
importance of the non-diagonal components of the permeability tensor should
be noted.
Narrow compositional range of enhanced magnetic softness in the samples
with negative magnetostriction constant and its strong dependence on the
conditions of processing are mainly determined by the magnetoelastic energy,
We' arising from the internal stresses, 0';. CT;. The estimated values of the
internal stresses in these amorphous microwires are of the order of 1000 MPa,
depending on the thickness of glass coating and metallic nucleus radius
aI., 1996; Chiriac and Ovari, 1997>.
(Velazquez et al., 1997). Such elevated internal
stresses give rise to a drastic change of the magnetoelastic energy, Kme~3/2
K me ~3/2
As 0';'
CT;, even for small changes of the magnetostriction constant, As' that is
when the composition sl slightly
ightly changes. Particularly, the high sensitivity of GMI
ratio on the applied stresses confirms the importance of the magnetoelastic
contribution.

5.6 Sensor Applications

Amorphous and nanocrystalline materials result to be very interesting for field


and stress-sensing elements. Generally, the Fe-rich amorphous alloys
exhibiting high magnetostriction values (As = 10- 5 ) and therefore many
magnetic parameters (i. e. , magnetic susceptibility, coercive field, etc.)
etc. ) are
extremely sensitive to the applied
appl ied stresses. On the other hand, Co-rich
amorphous materials with nearly-zero magnetostriction constant exhibit very
soft magnetic behavior with extremely low coercivity (of several Aim) and
high values of the initial permeability (up to 105 ) (Vazquez and Hernando,
1996). These peculiarities can be widely used in technological applications.
Magnetic sensors containing amorphous and nanocrystalline materials are
classified into six main groups:
1. Magnetometers using almost zero magnetostrictive amorphous and
nanocrystalline materials (flux sensors, current sensors, proximity sensors,
166 Arcady Zhukov and Julian Gonzalez

magnetometers) .
2. Stress sensors using high magnetostriction constant (i\ CA. ss =2 5-
= 2 x 10- 5-
4. 5 x 10-
4.5 5). These include:
10 - 5).
1) Sensors using bistable materials (pulse generators, encoders and
((1)
proximity sensors) .
( 2) Stress-magnetic sensors (torque sensors, force sensors, knock
sensors) .
(3) Ultrasonic sensors (distance sensors, frost sensors, pressure
sensors) .
3. Thermal sensors util izing the Curie temperature (T = - 50 - 250 'c
utilizing ).
'C).
4. Magnetic field sensors based on GMI effect of amorphous and
nanocrystalline
nanocrystall ine materials.
mater ials.
5. Sensors based on magnetoelectric effects (inverse Wiedemann and
Matteucci) .
6. Security and anti-thief sensors.
It seems that at the present moment the main attention is paid to the
development of magnetic sensors related to the use of metallic metall ic glasses in
fluxgate magnetometers and especially with the development of magnetic
sensors based on magnetic bistability (MB) and giant magneto-impedance
(GMI) effects (Vazquez and Hernando 1996; Mohri et al. , 2001; Vazquez,
(GM!)
2001). On the other hand, certain activity is also related with the use of
metallic glasses for security sensors (Sanchez et ai., al., 1988; Shin et al. ,
1992; Moron et al. , 1995), in delay lines (Hristoforou and Chiriac, 1995), in
cloth inductors (Matsuki and Murakami, 1985), etc.
Consequently main attention will be paid to the sensors based on the
magnetic bistability and GMI effect.
The latest advances in magnetic materials are based on the
miniaturization of modern magnetic materials. Therefore special attention will
be paid to the applications of tiny amorphous and nanocrystalline microwires
(in the order of 3 to 30 IJm~m in diameter) covered by an insulating glass coating
produced by the Taylor Ulitovski method.

5.6.1 Magnetometers Using Amorphous and Nanocrystalline Materials


Enhanced magnetic properties of amorphous and nanocrystalline materials
(mainly of nearly-zero magnetostrictive amorphous ribbons) gave rise to their
applications in fluxgate applications. One of the attractive features of metallic
glasses for use in fluxgate magnetometers is their low noise level at zero field.
Thus, Shirae (1984) was able to design a sensor of Vacquier type based on
(C067Fe3Si15BI5)C7
(C067Fe3Si15B15)C7 amorphous wire with a noise level as low as 2.5 pT
(RMS, 0.1 - 16 Hz). Ringcore fluxgate sensor with the noise level of 22 pT
O. 05 - 6 Hz was introduced by Narod based on CoFeSiB amorphous
RMS at 0.05
ribbon (Narod et al. , 1985). The main request for the material for the fluxgate
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 167

magnetometer is the sensitivity, determined by the coercivity. In order to


reduce the effect of the hysteresis, the coercivity should be of the order of
several IJT, or even in the nT range. Such sensibility has been achieved in
nearly-zero amorphous ribbons (Nielsen et ai., aI., 1990). Besides, in order to
improve the performance of the magnetometer, a "helical" magnetic
anisotropy has been induced by the annealing
anneal ing under torsion. In this way the
two-domain structure will set under the effect of the current flowing through the
ribbon. Such domain structure shows the sensitivity of around 15 AIT,
depending on the sample geometry, characteristics of the secondary coil, type
of the induced magnetic anisotropy and value of the current flowing through the
sample. Thus, the performance of fluxgate sensors based on 3 different
amorphous ribbons: quenched in air, in vacuum andlor stress annealed were
compared in (Nielsen
(N ielsen et al. , 1991).

5. 6. 2 Magnetoelastic Sensors: Sensors Based on Magnetic Bistability


Phenomena of the magnetic bistability characterized by the sharp voltage
peaks induced by abrupt change of the magnetic flux during the large
Barkhausen jump is very convenient for the design of magnetic sensors.
Magnetic sensors based on magnetic bistability were first introduced in the
1970s, when Wiegand wires (Wiegand, 1974) with rectangular hysteresis
loop with sharp 20 - 30 IJS appearing in the secondary coil mounted on the
sample subjected to the ac magnetic field were developed. Such sensors were
widely applicable in the automobile industry for the detection of the motions
and positions (Mohri et ai.,al., 1981). However, such Wiegand wires need
relatively large exciting field of the order of 4 kAlm for producing sharp
voltage peaks and special processing. Therefore, amorphous Fe-rich (Feso (Fe80 B20
20
or Fesl
Fe81 B 17 Si 2
2 )
) magnetostrictive ribbons subjected to the special stress
annealing were introduced in the beginning of the 1980s (Mohri et al. , 1981).
Stress annealing was performed in the samples of toroidal shape in order to
induce tensile-stressed and compressive-stressed layers in the ribbon. Sharp
voltage pulses with 20 - 50 IJS duration at field ampl
amplitude
itude of around 100 Aim
AI m at
frequency of O. 1 - 6 Hz were found in these amorphous ribbons. Such
magnetic bistability was initially used for the magnetometer and rotation speed
sensors (Mohri et al., 1981). Later, the Matteucci effect, appearing in the
twisted amorphous ribbon was also employed for the modified rotation speed
sensor (Mohri and Takeuchi, 1982). Like in the case of the sensor based on
magnetic bistability, sharp voltage pulses appearing between the sample
ends, were used as sensors (Mohri and Takeuchi, 1982). In this case the
secondary coil is not necessary.
Similarly, magnetic bistability (MB) has been observed in iron whiskers
(Heiden and Rogalla, 1982) and nearly-zero magnetostrictive C0 Co 70
70 Fe5 SilO B 15
amorphous ribbons (Gonzalez, 1996). In the case of amorphous Co-rich
168 Arcady Zhukov and Julian Gonzalez

ribbons a special heat treatment of the toroidal shaped sample has been used
(Ponomarev and Zhukov, 1984, 1985; Zhukov, 1993). Nevertheless, the use
of these materials is limited, due to intrinsic switching field fluctuations
observed in all these materials. Such switching field fluctuation results in
spontaneous experimental spread in the value of the switching field taking such
value in different re-magnetizing cycles (Heiden and Rogalla, 1982;
Ponomarev and Zhukov, 1984, 1985; Zhukov, 1993). These switching field
fluctuations can deteriorate the performance of the magnetic sensors based on
the magnetic a number of encoding combinations because of the overlapping of
the switching field values from different microwires and of the magnetoelastic
sensor making its work unreliable. A phenomenological model dealing with the
effect of thermal fluctuations on the remagnetization process have been
developed for the case of amorphous ribbons (Ponomarev and Zhukov, 1985;
Narod et a!.
al. , 1985; Nielsen et al. , 1990).
Throughout the 1980s and later, an increasing interest has been focused
on amorphous materials with acyl a cylindrical
indrical shape. The in-rotating-water
fabrication technique was widely employed for the production of around 120 IJm
amorphous wires. It was found that either as-cast or die-drawn then annealed
magnetostrictive amorphous wires exhibit the re-entrant flux reversal
characterized by large and single Barkhausen jump (Humphrey et al. , 1987a;
Gonzalez, 1996). Such particular behavior was explained by the particular
domain structure determined by the magnetoelastic anisotropy arising from the
internal stresses due to the fabrication process. It was supposed that rapidly
solidified core creates strong radial stresses in the outer shell. In this way,
almost single domain axially magnetized inner core with transverse easy axis
outer shell (radial domain structure in Fe-rich wires either circular for the
Co-rich) was assumed. Nearly-zero magnetostrictive compositions do not
show such behavior. A number of Fe- and Co-rich compositions (Fe-Si-B,
Fe-Co-Si-B, Fe-Cr-Si-B, Fe-Ni-Si-B, Co-Si-B) exhibited magnetic bistable
behavior were developed by the Unitika Ltd (Mohri et al., a!., 1990).
Consequently, a number of magnetic sensors based on magnetic bistabil ity
and/ or Matteucci effect of amorphous wires were developed (Humphrey,
1987b; Mohri et a!.al.,, 1990; Muzutani et a!.,
al., 1993; Ho and Yamasaki, 1996;
al. , 1998). As it was expected, sharp voltage peaks appearing in
Jahnes et a!.
the secondary coil or between the sample ends are commonly used for the
distance sensor (Mohri et a!.,
al., 1990), revolution counter and position sensor
(Muzutani et a!.,
al., 1993). On the other hand, such peculiar remagnetization
process with large Barkhausen discontinuity exhibited by amorphous wires has
been used for design of magnetic markers and tags (Humphrey, 1987b; Ho
and Yamasaki, 1996; Jahnes et al., aI., 1998). In these applications a
consequence of voltage peaks appearing under the application of the external
magnetic field is used. A general problem of such applications is the limited
number of combination for good identification due to the similarity of the
switching field values. In order to extend the switching field range, a coating
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 169

by the hard magnetic materials is recently used (Jahnes et al. , 1998).


The alternative technology or Taylor-Ulitovsky method for fabrication of
tiny fibers has been used recently to obtain glass coated amorphous
microwires with magnetic bistability (Zhukov et al., aI., 1995a). These
microwires consist of an amorphous metallic nucleus covered by an insulating
glass sheath. In this case, the range of the diameter of the metallic nucleus is
from about 1 to 20 IJm and the glass coating thickness is between 1- 10 IJm. It
was found that these amorphous microwires can also exhibit MB effect even for
samples with a length of a few millimeters and for nearly zero-magnetostrictive
alloys (Zhukov et al. , 1995a), and negative magnetostriction compositions do
not exhibit magnetic bistability. It is important to outline that even Fe-rich
glass coated microwires exhibit a wide range of the switching fields. Besides
such wide range of the switching field can be extended by the thermal
treatments (Arcas et ai., al., 1996). Particularly many experimental data
concerning spontaneous magnetic bistability of Fe-rich compositions were
explained mainly due to additional stresses arising from the outer glass cover
due to difference in thermal expansion coefficients with metallic
metall ic nucleus
(Zhukov, 2002).
Magnetic bistability with extended range of switching fields and high
stress sensitivity of magnetic parameters (switching, field) give rise to
various technological applications of tiny magnetic wires coated by glass
(Larin et al. , 1996).
One of the applications is based on a wide range of coercivity, which can
be obtained in microwires owing to its strong dependence on geometric
dimensions and heat treatment. It was realized in the method of magnetic
codification using magnetic tags (Larin et al. , 1996). The tag contains several
microwires with well-defined coercivities, all of them characterized by a
rectangular hysteresis loop. Once the magnetic tag is submitted to the ac
magnetic field, each particular microwire is remagnetized at a different
magnetic field giving rise to an electrical signal on a detecting system (Fig.
5.30). The extended range of switching fields obtained in Fe-rich microwires
gives a possibility to use a big number of combinations for magnetic
codification.
A magnetoelastic sensor of the level of liquid can be designed using the
stress dependence of the coercivity in nearly-zero magnetostriction CoMnSiB
microwires. Such a sensor essentially consists of a piece of the sample C
surrounded by primary and secondary coils (Fig. 5. 31) (Zhukov et al. ,
2000c). The sample is loaded with approximately 10 Grammies and therefore
exhibits
exhib its a flat hysteresis loop. The weight is attached to the bottom of the
sample. When a liquid arrives to cover the weight as a consequence of the
actual stress, affecting the sample, decreases giving rise to the appearance of
the rectangular hysteresis loop. The principle of the sensor's work is based on
the change of the voltage of the secondary coil, which increases drastically
when the hysteresis loop of the sample C became rectangular after essential
170 Arcady Zhukov and Julian Gonzalez

Al
M

-
H c2
\~
H el If
H
- H c3
r-

00000001001000001000000000100100000010
()OI~Oi)(IOjOOlOOOOOIOOOU(I()uUOI()OIOOUOOUIO

Digital m ....r r I. . ..-rTTlr..-TTlrTTTTl-.-. .,.,-TTT.....-1


Di gi tal f-r-r"rrI..,...,..-rTT1c.-TT1rTTTT1c.-,....,,-rrr...-,
codes

Figure 5. 30 Schematic representation of the encoding system based on magnetic


5.30
bistability of the microwires.

16

J~::::::-=~
L--l~cI:-=~ Sample
',.-_.-----< 14
t:::c-=3 Input
~

E 12
:;:
TI;-L-<
1-
Output
Output
I=! *::::;
10
.--- __ Load
load
I - -" Level of the
1
i liquid 8
L-- -'- -L-- -"-_

a 100 200 300


c; (MPa)
<Y

5. 31 Schematic representation of the magnetoelastic sensor based on stress


Figure S.
dependence of the switching field.

decreasing of the stress. Simple circuit including the amplification of the signal
and alarm set was used to detect changes of voltage in the secondary coil
under the change of tensile stresses owing the floating effect of the weight on
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 171

the liquid.
liquid
The observed magnetic response on the external variable stresses can be
used in many appl ications to detect different temporal changes of stresses,
vibrations, etc. As an example we introduce a "magnetoelastic pen" which
vibrations.
can be used for identification of signatures (Zhukov et al.
al , 1998). It is known
that the signature of each person can be represented by a typical series of the
stresses. The sequence and strength of those stresses are a characteristic
feature of any personal signature. So.
So, temporal changes of the stresses while
signing can be used for the identification of a signature itself. We have used
this behavior and designed a set-up consisting of a ferromagnetic bistable
amorphous wire with positive magnetostriction, a miniaturized secondary coil
and a simple mechanical system inside the pen containing a spring, spring. which
transfers the applied
appl ied stresses to the ferromagnetic wire (Fig. 5. 32a). The
stresses. corresponding to a signature is
resulting temporal dependence of the stresses,
reproducible for each individual person. The main characteristics of this
dependence are time of signature, sign and sequence of the detected peaks.
Examples of two magnetoelastic signatures corresponding to the same person
are shown in the Fig. 5. 32b.

~li---~=S
Spring ~- rm1
Pick up coil
(a)

1st
15t signature

;;-
>
'0
c o 4 6
Time(s)
Time (s)

2nd signature

Time(s)
(b)

Figure 5.32 Schematic representation of the magnetoelastic pen (a) and two
magnetoelastic signatures (b)
(b).
172 Arcady Zhukov and Julian Gonzalez

5.6.3 Sensors Based on Giant Magneto-Impedance Effect


Since the discovery of the magneto-impedance effect in 1994 (Panina and
Mohri 1994; Beach and Berkowicz, 1994), a number of researches and
developments in this direction have been widely performed by different groups
(Vazquez et al.,aI., 1997). Thus, a CMOC IC circuitry with pulse current
operation was established for the Ml MI sensor (Kanno et al. , 1997). The stress-
impedance (SI) and torsion impedance (TI) effects showing high sensitivity of
the impedance to the applied stress with a strange gauge factor of 2000 - 4000
have been found in amorphous wires (Shen et al. , 1997; Blanco et al. , 1999a
and 1999b).
As a result, the GMI and SI Sl sensors with the CMOC IC circuitry with
advantageous features comparing with conventional magnetic sensors have
been developed by Aichi Steel Co (Mohri et al. , 2001). Main applications
appl ications are
related with the detection of magnetic fields, small weights and vibrations,
and such branches of the industry as the car industry and medicine are main
consumers of these sensors. Among the almost 100 applications, the following
sensors are currently in progress: portable digital display of the terrestrial
magnetic field, brain tumor sensor, secondary current sensor for the induction
motor control, car passing measurement and recording disk, finger-tip blood
vessel pulsation, mechanoencephalogram sensor, etc. The sensor's circuitry,
performance and comparative characteristics with the traditional sensors are
presented in (Mohri et aI., al., 2001). Similarly, the other sensors based on
stress sensitivity of the GMI effect for domestic use have been developed by
other groups (Cobeno et al., al. , 2001). As an example the air flux sensor is
described below.
Axial field dependencies of GMI ratio, D.Z I Z, measured in glass coated
amorphous C0 68 5 Mnu SilO B 15 microwire at 10 MHz and I = 1.5 mA and at
COGB
different tensile stresses, cr, a, are presented in the Fig. 5. 33. The tensile
stresses ranging between 0 and 132 MPa correspond to the mechanical load of
0- 2. 9 gram. In the unstressed state there is a maximum in (D.Z I Z) m for a
value of around H m = = 120 Aim of axial applied field. This maximum displaces
(D.ZI Z)m first increases up to
towards larger applied fields and the value of (D.ZIZ)m
130 % and then decreases with cr. a.
This high sensitivity of GMI ratio to quite small values of mechanical load
makes this stress sensitive GMI effect very interesting for practical purposes.
Figure 5.33 shows a sensitive magnetoelastic device based on GMI exhibited
by this microwire with the following principle of work. Such as it was
mentioned by Vazquez and Zhukov (1996), the microwire has rather high dc
electrical resistance per length owing to its tiny dimensions. Therefore, large
changes of the magneto impedance could induce rather large changes of the ac
voltage under mechanical loading. Indeed, under the effect of tensile stresses,
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 173

150
ISO r
120 ~~4
~Ol5d~
t~ 1 90 0 D~~~l!:
~ 60 ~I ~ WithOlJt_st~e~~~~~",
N
~ D 00
.P=O.72g 11
Air flux 30
30 tc 0 0
.0
!~o L'>P=I.44 ~
TP=2.16g
0- ~e OP~2.88g
o .L---=2--'-OCC"0--"'------:-40""'0'---

.
__.-- wI i c 1'0\\-' i re
_Microwire fI(A/m)

fide field \J
(A'j ac [)()\\
power er source

10
lOr
..... ..
'." '. -.
'----+~l--Sail
'------+~l-- Sail

):
>
~
"
Of]
')1)
'El 8
3~
f
9
'. ..'.
..'. ....
'., '...- ..
:l
u
'.)

'"'"' 7

6
0
peg)
Pig)
2
-

3
.
'.
Figure 5.33 Schematic representation of the magnetoelastic sensor based on stress
dependence of GMI effect (a) and its calibration curve (b).

the GMI ratio at fixed applied field, H, decreases giving rise to a significant
change of the ac voltage between the sample's ends.
A dc magnetic field corresponding to the maximum ratio (!:.Z/ Z) m without
(I:::.Z / Z)m
stresses has been applied by means of a small solenoid of 7 cm in length and
1 cm in diameter. In the case of a driving current with the frequency 10 MHz
ampl itude 1.5 mA suppl
and amplitude ied by an ac current generator, the change of ac
supplied
voltage (peak to peak) between the ends of the sample is of around 3.5 V for
small mechanical loads attached at the bottom of the microwire (see
calibration curve in the Fig. 5. 33). This huge change presented at the output
permits us to think of the successful use of this kind of sensor in different
technological applications related with the detection of alternative mechanical
stress.
In order to obtain an ac high frequency input, a source of ac current has
been introduced. Such source consists of the power ampl ifier based on
amplifier
transistor T1 (KT 603 A) and the microscheme K561 K561LA7.
LA7. Such microscheme
contains 3 logical elements (LE 1, LE2, LE3). The resistor R 1 needs to
establ ish a steady autooscillation, with a capacity C
establish 0,1 for the positive
feedback. The frequency, f, of the oscillations depends on the value of such
capacity 0,
C 1 ..
174 Arcady Zhukov and Julian Gonzalez

References
Alben, R. , J. J. Becker and M. C. Chi. J. Appl. Phys. 49: 1653 (1978a)
Alben, R.A., J.I.J.1. BudnickandG.S.
Budnick and G.S. Cargill (III). In: J.J. GilmanandH.J.
Gilman and H.J.
Leamyeds.
Leamy eds. Metallic
Me tal/ ic Glasses. American
Amer ican Society of Metals, Metals Park,
Ohio, p394 (1978b)
AIIia, P. , G. P. Soardo and F. Vinai. Solid State Communications 24: 517
(1977)
( 1977)
Aragoneses, P. , A. Zhukov, J. Gonzalez, J. M. Blanco and L. Dominguez.
Sensors and Actuators A 81: 86 (2000)
Aranda, G. R. , C. Miguel, P. Garcia-Tello and J. Gonzalez. J. Appl. Phys.
89: 6422 (2001)
Areas, J., C. G6mez-Polo,
Gomez-Polo, A. Zhukov, M. Vazquez, V. Larin and A.
Hernando. Nanostructured Materials 7: 823 (1996)
Argon, A.A.S.S. andH.Y.
and H. Y. Kuo. J. Non-Crystalline Solids. 37,241
37, 241 (1980)
Atkinson, D. andP.T. Squire. IEEE Trans. Magn. Mag-33: 3364 (1997)
Barandiaran, J. M. , A. Hernando, V. Madurga, O. V. Nielsen, M. Vazquez
and M.M . Vazquez-L6pez.
Vazquez-Lopez. Phys.Rev.
Phys. Rev. B 35: 5066 (1987)
Beach, R. S. and A. E. Berkowicz. Appl. Phys. Lett. 64: 3652 (1994)
Blanco, J. M., P. G. Barb6n, Barbon, A. R. Pierna and J. Gonzalez. J. Non-
Crystall ine Sol
Crystalline ids 136: 91 (1991)
Solids
Blanco, J. M. , P. Aragoneses, J. Gonzalez, A. Hernando, M. Vazquez, C.
G6mez-Polo,
Gomez-Polo, J. M. Barandiaran, P. T. Squire and M. R. J. Gibbs. In: L.
Lanotte ed. Magnetoelastic Effects and Applications. ElsevierSci. Publ. B.
V. , Amsterdam, 253 (1993)
Blanco, J. M., L. Dominguez, P. Aragoneses and J. Gonzalez. J. Magn.
Magn.Mat.
Magn. Mat. 186: 135 (1998)
Blanco, J. M. , A. Zhukov and J. Gonzalez. J. Magn. Magn. Mat. 196 -197:
377 (1999a)
Blanco, J. M. , A. Zhukov and J. Gonzalez. J. Phys. D: Appl. Phys. 32: 3140
((1999b)
1999b)
Blanco, J. M. , A. Zhukov and J. Gonzalez. J. Appl.App!. Phys. 87: 4813 (2000)
Blanco, J. M. , A. Zhukov, A. P. Chen, A. F. Cobeno, A. Chizhik and J.
Gonzalez. J. Phys. D: App!.Appl. Phys. 34: L31 (2001)
Callen E. and H.B.
H. B. Callen. Phys. Rev. A 139: 455 (1965)
Phys.Rev.
Chen, D.X. andL.C.
and L.C. Tai. J. Magn. Magn. Mat. 50: 329 (1985)
J.Magn.Magn.Mat.
Chen, D.X. J.Appl.
J.Appl.Phys.
Phys. 61: 3781 (1987)
Chen, D. X. , C. Gomez-Polo and M. Vazquez. J. Magn. Magn. Mat. 124: 262
(1993)
( 1993)
Chen, D. X. , N. M. Dempsey, M. Vazquez and A. Hernando. IEEE Trans.
Magn. Mag-31: 781 (1995)
Chiriac, H. , M. Knobel and T. Ovari. Materials Science Forum 302 302-303:
- 303: 23
(1999)
Amorphous and Nanocrystalline 50ft
Soft Magnetic Materials: Tailoring... 175

Chiriac, H. andandT.A.
T. A. Ovari. Progress in Material Science. 40: 333 (1997)
Ciureanu, P. , P. Rudkowski, G. Rudkowska, D. Menard, M. Britel, J. F.
Currie, J. O. Strom-Olsen and A. Yelon. J. Appl. Phys. 79: 5136 (1996)
Cobefio, A. F., A. Zhukov, A. R. de Arellano - Lopez, F. Elias. J. M.
Cobeno,
Blanco, V. Larin and J. Gonzalez. J. Mat. Res. 14: 3775 (1999)
Cobefio, F. A. , A. Zhukov, J. M. Blanco, V. Larin, J. Gonzalez. Sensors
Cobeno,
and Actuators A 91: 95 (2001)
Corb, R. W.,W. , R. C. 0' Handley, J. Megusar and M. J. Grant. Phys. Rev.
Lett. 51: 1386 (1983)
De Lacheisserie, Du Tremolet E. and R. Krishnan. J. Appl. Phys. 55: 2461
(1984)
( 1984)
De Lacheisserie, Du Tremolet E. and J. Gonzalez. J. de Physique 50: 949
(1989)
( 1989)
Duwez, P., R.J. WiliiamsandK.
WilliamsandK. Klement. J.Appl.Phys.
J.AppI.Phys. 31: 1136 (1966a)
Duwez, P. In: H. Reiss ed. Progress in Solid State Chemistry of AI/oy
Phases. vol.3.
vol. 3. Pergamon Press, Oxford, p.377 (1966b)
Egami, T. and P.J. Flanders. IEEE Trans. Magn. Mag-H. 220 (1976)
Fi:ihnle, M. , J. Furthmuller, R. Pawellek and T. Beuerle. Appl. Phys. Lett.
59: 204 (1991)
Frost, M., I. Todd, H.A. Davies, M.R.J. GibbsandR.V.
Gibbs and R. V. Mayor. J.Magn.
Magn. Mat. 203: 85 (1999)
Fujimori, H.,
H. , T. Masumoto, Y. Obi and M. Kikuchi. Jap. J. Appl. Phys. 13:
1889 (1974)
Fujimori, H, K.1.
K. I. Arai, H. Shirae, H. Saito, T. Masumoto and N. Tsuya.
Jap.J.Appl.
Jap.J.AppI.Phys.
Phys. 15: 705 (1975)
Gambino, R.L., G.L. Saran. J.Appl.Phys.
J.AppI.Phys. 50: 7671 (1979)
Garcia, P. M. J. , E. Pina, A. P. Zhukov, V. Larin, P. Marin, M. Vazquez
and A. Hernando. Sensors and Actuators A 81: 226 (2000)
Gonzalez, J., M. Vazquez, J. M. Barandiaran, V. Madurga and A.
Hernando. J. Magn. Magn. Mat. 68: 151 (1987)
Kulakowski.i. J. Magn. Magn. Mat. 82: 94 (1989)
Gonzalez, J. and K. Kulakowsk
Gonzalez, J. and E. Du Tremolet De Lacheisserie. J. Magn. Magn. Mat. 78:
237 (1989)
Gonzalez, J. , J. M. Barandiaran, M. Vazquez and A. Hernando. Anales de
Fisica (B) 86: 184 (1990a)
Gonzalez, J. , J. M. Blanco, I. Telleria, J. M. Barandiaran, M. Vazquez, A.
Hernando and A. A.R.
R. Pierna. J. Magn. Magn. Mat. 83: 168 (1990b)
Gonzalez, J. J. Magn. Magn. Mat. 87: 111 (1990c)
Gonzalez, J., J. M. Blanco, J. M. Barandiaran, M. Vazquez and A.
Hernando. In: W. Gorzkowski, M. Gutowski, H. K. Lachowicz and H.
Szymczak eds. Physics of Magnetic Materials. World Scientific,
Singapore, 354 (1991)
Gonzalez, J and J. M. Blanco. J. Mat. Res. 7: 1602 (1992)
Gonzalez, J.,J. , E. Irurieta, P. Aragoneses, J. M. Blanco, I. Ibarrondo, M.
176
'76 Arcady Zhukov and Julian Gonzalez

Vazquez, J. M. Gonzalez and F. Cebollada. IEEE Trans. Magn. Mag-30:


1015 (1994)
Gonzalez, J. J. Appl. Phys. 79: 376 (1996)
Gonzalez, J. M. , N. Murillo, J. Gonzalez, J. M. Blanco and J. Echeverria.
J. Mat. Res. 11: 512 (1996)
Gonzalez, J. M. , C. de Julian, F. Cebollada, A. K. Giri and J. Gonzalez. J.
Appl. Phys. 81: 4658 (1997)
Graham, C.D., T. Egami. In: R.A. Huggins, R.H. BubeandR.W. Roberts
ed. Annual Review of Material Science, Annual Reviews Inc. Palo Alto,
CA, 8: 423 (1978)
Gubanov, A.I. Fizika 2: 502 (1960)
Guntherodt, H. -J. Glassy Metals 1/,II, Springer Verlag, Berlin
Berl in (1978)
Gutierrez, J.,
J. , P. Gorria, J. M. Barandiaran and A. Garcia-Arribas. In: A.
Hernando and M. Vazquez eds. Nanostructured and Non-Crystalline Non-Crystal/ine
Materials. World Sci. Publ. , Singapore, p. 500 (1995)
Handrich, K. Phys. Stat. Sol. 32: K55 (1969)
Phys.StaLSol.
Hernando, A. , M. Vazquez, V. Madurga and H KronmUiler.
KronmOlier. J. Magn. Magn.
Mat. 37 (1983)
Hernando, A., V. Madurga, C. Nunez de Villavicencio and M. Vazquez.
Appl. Phys. Lett. 4S:45: 802 (1984)
Heiden, C. and H. Rogalla. J. Magn. Magn. Mat. 26: 275 (1982)
Hernando, A. , V. Madurga, J. M. Barandiaran and O. V. Nielsen. Solid State
Communications. S4: 54: 1059 (1985)
Hernando, A., C. Gomez-Polo, E. Pulido, G. Rivero, M. Vazquez, A.
Giarca-Escorial and J.M. Barandiaran. Phys.Rev. B 42: 6471 (1990)
Hernando, A. , J. Gonzalez, J. M. Blanco, M. Vazquez, J. M. Barandiaran,
G. Rivero and E. Ascasibar. Phys. Rev. B 46: 3401 (1992)
Hernando, A. and M. Vazquez. In: H. H. Liebermann ed. Rapidly Solidified
Alloys. Marcel Dekker, New York, p.553 (1993)
Hernando, A., M. Vazquez, T. Kulik. and C. Prados. Phys. Rev. B 51:
3581 (1995)
Hernando, A., P. Marin, M. Vazquez, J. M. Barandiaran and G. Herzer.
Phys. Rev. B S8:58: 366 (1998)
Herzer, G. IEEE Trans. Magn. Mag-26, 1397 (1990a)
Herzer, G. Anales de Fisica B 86: 157 (1990b)
Herzer, G. Physica Scripta T 49: 307 (1993)
Herzer, G. J. Magn. Magn. Mat. 133: 1 (1994a)
Herzer, G. IEEE Trans. Magn. Mag-30: 1800 (1994b)
Herzer, G. Scr. Metall. Mater. 33: 1713 (1995a)
Herzer, G. In: A. Hernando and M. Vazquez eds. Nanostructured and Non-
Crystalline Materials. World Scientific, Singapore, p.449
p. 449 (1995b)
Hilzinger, H. R. In: T. Masumoto and K. Suzuki eds. Proceedings of 4th
Intern. Conference on Rapidly Quenched Metals. Jap. Inst. Metals,
Sendai, p. 781 (1982)
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 177

Ho, W. and J. Yamasaki. US patent 5.519.379, May 21 (1996)


Hofmann, B. and H. Kronmuller. J. J.Magn.Magn.Mat.
Magn. Magn. Mat. 152: 91 (1996)
Hristoforou, E. and H. Chiriaco In: M. Vazquez and A. Hernando eds.
Nanostructured and Non-crystalline Solids. World Scientific Co, Singapore,
p.505 (1995)
Humphrey, F. B. , K. Mohri, J. Yamasaki, H. Kawamura, R. Malmhi:ill, Malmhall, I.
Ogasawara. In: A. Hernando, V. Madurga, M. C. Sanchez-Trujillo and M.
Vazquez eds. Magnetic Properties of Amorphous Metals. Elsevier Science
Publ.,
Pub\. , Amsterdam, p. 110 (1987a)
Humphrey, F. B. US patent 4.660.025, Apr. 21 (1987b)
Jagielinski, T. , K. I. Arai, N. Tsuya, S. Ohnuma and T. Masumoto. IEEE
Trans. Magn. Mag-13: 1553 (1977)
Jahnes, C.,
C. , R.J.
R. J. Gambino, M. Paunovic, A. G. Schroff, R.J. R. J. von Guftel.
Gufte\.
US patent 5.729.201,
5. 729. 201, Mar. 17 (1998)
Progress Phys. 36: 1425 (1973)
Jones H. Rep. Progres$
Kanno, T., K. Mohri, T. Yagi, T. Uchiyama, L. P. Shen. IEEE Trans.
Magn. Mag-22: 3358 (1997)
Kikuchi, M. , H Fujimori, Y. Obi and T. Masumoto. Jap. J. App\. Appl. Phys. 14:
1077 (1975)
Klement, K.,K. , R.H.
R. H. Wilens and P. Duwez. Nature 187: 869 (1970)
Knobel, M. , C. Gomez-Polo and M. Vazquez. J. Magn. Magn. Mat. 160: 243
((1996)
1996)
Kraus, L.,
L. , V. Haslar, O. Heczko and K. Zaveta. J. Magn. Magn. Mat. 157/
158: 151 (1996)
Kronmuller, H., M. Fahnle, M. Domann, H. Grimm, R. Grimm and B.
Groger. J. Magn. Magn. Mat. 13: 53 (1979)
Kronmuller, H. J. Appl. Phys. 52: 1859 (1981)
J.App\.Phys.
Lachowicz, H. K. and A. Siemko. IEEE Trans. Magn. Mag-25. 3605 (1989)
Lachowicz, H. K. , A. Neuweiler, F. Poplawski and E. Dynowska. J. Magn.
Magn. Mat. 173: 287 (1997)
Larin, V., A. Torcunov, S. Baranov, M. Vazquez, A. Zhukov and A.
Hernando. Method of magnetic codification and marking of the objects.
Patent (Spain) No P 9601993 (1996)
Lim, S. H. , W. K. Pi, T. H. Noh, H. J. Kim and I.\. K. Kang. J. Appl. Phys.
73: 6591 (1993)
Luborsky, F.E., J.J. BeckerandR.O. McCary. IEEE Trans. Magn. Mag-ll: Mag-H:
1644 (1975)
Luborsky, F. F.E.
E. and
andJ.L.Walker.
J. L. Walker. IEEE Trans. Magn. Mag-13: 953 (1977)
Luborsky, F. E. and H. H. Liebermann. Mater. Sci. Eng. 49: 257 (1981 a)
Luborsky, F.E.,
F. E. , S.C.
S. C. Huang and H.C.
H. C. Fiedler. IEEE Trans. Magn. Mag-17:
3463 (1981b)
( 1981 b)
Luborsky, F. E. In: F. Luborsky ed. Amorphous Metallic Alloys. Butterworth
& CoPublishers Ltd (1983)
Machado, F. L. , C. S. Martins and B. M. Razebde. Phys. Rev. B 51: 3926
178 Arcady Zhukov and Julian Gonzalez

(1995)
Pan ina and D. J. Mapps. Phys. Rev. B 63:
Makhnovskiy, D. P., L. V. Panina
144,424 (2001)
Makhotkin, V. E. , B. P. Shurukhin, V. A. Lopatin, P. Yu. Marchakov and Yu.
K. Levin. Sensors and Actuators A 25 - 27: 759 (1991)
Marin, P., M. Vazquez, L. Pascual, D. Negri, F. Leccabue, B.E. Watts,
H.A.
H. A. Davies and A. Hernando. Mater. Sci. Forum 235: 743 (1997>
(1997)
Matsuki, H., K. Murakami. IEEE Trans. Magn. Mag-2l:
Mag-21: 1738 (1985)
Miguel, C. , N. Murillo and J. Gonzalez. J. Appl. Phys. 88: 6623 (2000)
Miyazaki, T. andM.
and M. Takahashi. Jap.J.Appl.Phys. 17: 1755 (1978)
Miroshnichenko, I.S. andl.V. Salli. Ind. Lab. 25:1463 (in English); Zav.
Lab. 25: 1398 (1959)
Mizoguchi, T.,
T. , K. YamaguchiandH. H. O. Hooper and A.M.
Yamaguchi and H. Miyajima. In: H.O. A. M.
de Graaf Eds. Amorphous Magnetism. Plenum Press, New York, p. 325
((1973)
1973)
Mohri, K., B. Takeuchi and T. Fujimoto. IEEE Trans. Magn. Mag-17: 3370
( 1981)
Mohri, K. and S. Takeuchi. J.Appl.Phys.
J. Appl. Phys. 53: 8386 (1982)
Mohri, K. IEEE Trans. Magn. Mag-20: 942 (1984)
Mohri, K. , F. B. Humphrey, K. Kawashima, K. Kimura and M. Muzutani.
IEEE Trans. Magn. Mag-26. 1789 (1990)
Mohri, K. , T. Uchiyama, L. P. Shen, C. M. Cai, L. V. Panina. Sensors and
Actuators A 91: 85 (2001)
Moron, C., C. Aroca, M. C. Sanchez, A. Garcia, E. Lopez and P.
Sanchez. IEEE Trans. Magn. Mag-31: 906 (1995)
Muller, N. and N. Mattern. J. Magn. Magn. Mat. 136: 79 (1994)
Murillo, N. , J. M. Blanco, J. Gonzalez. and M. Vazquez. J. Appl. Phys. 74:
3323 (1993)
Murillo, N. and J. Gonzalez. J. Magn. Magn. Mat. 218: 53 (2000)
Muzutani, M. H. Katoh, L. V. Panina,
Pan ina , K. Mohri and F. B. Humphrey. IEEE
Trans. Magn. Mag-29: 3174 (1993)
Narod, B. B. , J. R. Bennest, J. O. Strom- Olsen F. Nezil and R. A. Dunlap.
Can. J. Phys. 63: 1468 (1985)
Neel, M. L. J. Phys. Radium 15: 225 (1954)
J. Magn.Magn. Mat. 22: 21 (1980a)
Nielsen, O.V. andH.J.V. Nielsen. J.Magn.Magn.Mat.
Nielsen, O. V. and H. J. V. Nielsen. Solid State Communications 35: 281
( 1980b)
Nielsen, O.V. J.Magn.Magn.Mat. 36: 73 (1983)
Nielsen, O. V., A. Hernando, V. Madurga and J. M. Gonzalez. J. Magn.
Magn. Mat. 46: 125 (1985)
Nielsen, O. V., B. Hernando, J. R. Petersen and F F. Primdahl. J. Magn.
Magn. Mat. 83: 405 (1990)
Nielsen, O. V. , J. R. Petersen, A. Fernandez, B. Hernando, P. Spisak, F.
Primdahl and N. Moser. Meas Sci. Techn. 2: 435 (1991)
Amorphous and NanocrystaJline
Nanocrystalline Soft Magnetic Materials: Tailoring... 179

Nielsen, O. V. , J. R. Petersen and G. Herzer. IEEE Trans. Magn. Mag-30:


1042 (1994)
Obi, Y., H. Fujimori and H. Satio. Jap.J.Appl.Phys.
FujimoriandH. Jap.J.AppI.Phys. 23: 281 (1976)
Ogasawara, I. and S. Ueno. IEEE Trans. Magn. Mag-31: 1219 (1995)
O'Handley, R. C. IEEE Trans. Magn. Mag-ll: 206 (1975)
0' Handley, R. C. , R. Hasegawa, R. Ray and C. -P- Chou. Appl. Phys. Lett.
29: 330 (1976)
O'Handley, R. C. Phys.
R.C. Phys.Rev.
Rev. B8 18: 930 (1978)
O'Handley, R.CandM.D.
R.C and M.D. Sullivan. J.AppI.Phys. 52: 1841 (1981)
(1981>
Panina, L.V. and K. Mohri. Appl.Phys.Lett. 65: 1189 (1994)
Ponomarev, B. K. A. Zhukov. Sov. Phys. Sol id State 26: 2974 (1984)
Ponomarev, B. K. and A. P. Zhukov. Sov. Phys. Sol id state 27: 444 (1985)
Reininger, T., H. Kronmuller, C. Gomez-Polo and M. Vazquez. J. Appl.
Phys. 73: 5357 (1993)
Riehemann, W. and E. Nembach. J. Appl. Phys. 55: 1081 (1984)
Sanchez P. , E. Lopez, C. Aroca, M. C. Sanchez. IEEE Trans. Magn. Mag-
24: 1981 (1988)
Severino, A. M. , C. Gomez-Polo, P. Marin and M. Vazquez. J. Magn. Magn.
Mat. 103: 117 (1992)
Shen, L. P. , T. Uchiyama, K. Mohri, E. Kita, K ita, K. Bushuda. IEEE Trans.
Magn. Mag-33: 3355 (1997)
Shin, K. H., C. D. Graham, P. Y. Zhou. IEEE Trans. Magn. Mag-28: 2772
(1992)
Shirae, K. IEEE Trans. Magn. Mag-20: 1299 (1984)
Sinnecker, J. P. , E. H. C. P. Sinnecker, A. Zhukov, J. M. Garcia-Beneytez,
M. J. Garcia Prieto and M. Vazquez. J. Phys. IV France 8: Pr2-225
((1998)
1998)
Sixtus, K.J. andL.
and L. Tonks. Phys.Rev. 42: 419 (1932)
Slawska-Waniewska, A. , R. Zuberek and P. Nowicki. J. Magn. Magn. Mat.
157/158: 147 (1996a)
Slawska-Waniewska, A., R. Zuberek, J. Gonzalez and H. Szymczak.
Mater. Sci. Eng. A (Supplement), 220 (1996b)
Suzuki, K., G. Herzer and M. Cadogan. J. Magn. Magn. Mat. 177: 949
(1998)
Szymczak, H. J. Magn. Magn. Mat. 67: 227 (1987)
Takahashi, M., T. Suzuki and Y. Miyazaki. Jap. J. Appl. Phys. 16: 521
( 1977)
Tate, B.J., B.S. Parmere, I. Todd, H.A. Davies, M.R.J. GibbsandR.V.
Mayor, J. Appl. Phys. 83: 6335 (1998)
Tejedor, M.,M. , B. Hernando, M.L.
M. L. Sanchez, V.M.
V. M. Prida and M. Vazquez. J.
Phys.D: Appl.Phys. 31: 3331 (1999)
Twarowski, K. , M. Kurzminski, A. Slawska-Waniewska, H. Lachowicz and
G. Herzer. J. Magn. Magn.Mat. 150: 85 (1995)
J.Magn.Magn.Mat.
Usov, N.A., A.S. Antonov and A.N. Lagar'kov. J.Magn.Magn.Mat. 185:
180 Arcady Zhukov and Julian Gonzalez

159 (1998)
Varga, L. K. , L. Novak and F. Mazaleyrat. J. Magn. Magn. Mat. 210: L25
(2000)
Vazquez, M. , A. Hernando and O. V. Nielsen. J. Magn. Magn. Mat. 61: 390
(1986)
Vazquez, M. , E. Ascasibar, A. Hernando and O. V. Nielsen. J. Magn. Magn.
Mat. 66: 37 (1987a)
Vazquez, M., J. Gonzalez, V. Madurga, J. M. Barandiaran, A. Hernando
and O. V. Nielsen. In: A. Hernando, V. Madurga, M. C. Sanchez-Trujillo
and M. Vazquez eds. Magnetic Properties of Amorphous Metals. Elsevier
Sci. Publ. B. V. , Amsterdam, p.324
p. 324 (1987b)
Vazquez, M. , P. Marin, H. A. Davies and A. O. Olofinjana. Appl. Phys. Lett.
64: 3184 (1994)
Vazquez, M., A. Zhukov. J.Magn.Magn.Mat. 160: 223 (1996)
Vazquez, M. and A. Hernando. J. Phys. D: Appl. Phys. 29: 939 (1996)
Vazquez, M., M. Knobel, M. L. Sanchez, R. Valenzuela and A. Zhukov.
Sensors and Actuators A 59: 20 (1997)
Vazquez, M., A. Zhukov, P. Aragoneses, J. Areas, P. Marin and A.
Hernando. IEEE Trans. Magn. Mag-34. 724 (1998)
Vazquez, M. Physica B 299: 302 (2001)
Velazquez, J. , M. Vazquez and A. Zhukov. J. Mat. Res. 11: 2499 (1996)
Wiegand, J.R. US Patent 3,820,090 (1974)
Wiegand, J.R. US Patent 4,247,601 (1981)
(1981>
Wolfarth, E. P. Ferromagnetic Materials. North Holland Publ. Co,
Amsterdam (1980)
Yamamoto, T. Development of Sendust, Chiba, Kamiyama Komiyama Pr., p. 27
(1980)
Yoshinaga, Y., S. Furukawa and K. Mohri. IEEE Trans. Magn. 35: 3613
(1999)
Yoshizawa, Y., S. OgumaandK. Yamauchi. J.Appl.Phys.
J.AppI.Phys. 64: 6044 (1988)
Zhukov, A. Materials and Design 14: 299 (1993)
Zhukov, A. , M. Vazquez, J. Velazquez, H. Chiriac and V. Larin. J. Magn.
Magn. Mat. 151: 132 (1995a)
Magn.Mat.
Zhukov, A, J. Velazquez, E. Navarro and M. L. Sanchez. In: M. Vazquez
and A. Hernando eds. Nanostructured and Non-crystalline Solids. World
Scientific, Singapore, p. 542 (1995b)
Zhukov, A. , J. M. Garcia-Beneytez and M. Vazquez. J. de Physique IV 8:
Pr2-763( 1998)
Zhukov, A. , E. Sinnecker, D. Paramo, F. Guerrero, V. Larin, J. Gonzalez
and M. Vazquez. J. Appl. Phys. 85: 4482 (1999)
Zhukov, A., J. Gonzalez, J. M. Blanco, M. J. Prieto, E. Pina and M.
Vazquez. J. Appl. Phys. 87: 1402 (2000a)
Zhukov, A.,
A. , J. Gonzalez, J. M. Blanco, M. Vazquez and V. Larin. J. Mat.
Res. 15: 210 (2000b)
Amorphous and Nanocrystalline Soft Magnetic Materials: Tailoring... 181

Zhukov, A. , J. Gonzalez, J. M. Blanco, P. Aragoneses and L. Dominguez.


Sensors and Actuators A 81: 129 (2000c)
Zhukov, A. Appl. Phys. Lett. 78: 3106 (2001)
Zhukov, A. J. Magn. Magn. Mat. 249/1-2: 318 (2002)
Zhukova, V., A. F. Cobefio, A. Zhukov, J. M. Blanco, V. Larin and J.
Gonzalez. Nanostructured Materuials 11 (8): 1319 (1999)
Zhukova, V., A. F. Cobefio, E. Pina, A. Zhukov, J. M. Blanco, L.
Dominguez, V. Larin and J. Gonzalez. J. Magn. Magn. Mat. 215 - 216216::
322 (2000)
Zhukova, V. A., A. F. Cobefio, A. Zhukov, J. M. Blanco, S. Puerta, J.
Gonzalez and M. Vazquez. J. Non-Crystalline Solids 287: 31 (2001)
Zhukova, V. A., A. B. Chizhik, J. Gonzalez, D. P. Makhnovskiy, L. V.
Panina, D. J. Mapps and A. Zhukov. J. Magn. Magn. Mat 249/1-2: 318
(2002)
Zuberek, R., H. Szymczak, R. Krishnan, C. Sela and M. Kaabouchi. J.
Magn. Magn. Mat. 121 :510 (1993)
121:510

The authors are grateful to Ms. V. Zhukova and Dr. A. Chizhik for the help with the chapter
preparation and wish to acknowledge the contribution of Dr. J. M. Blanco.
6 Nanogranular Layered Magnetic Films

G. N. Kakazei.
Kakazei, Yu. G. Pogorelov.
Pogorelov, J. B. Sousa.
Sousa, J. M. Santos.
Santos,
S. Cardoso and P. P. Freitas

6. 1 Introduction

Nanogranular layered magnetic films are a new class of magnetoresistive (MR)


systems, consisting of well separated layers of closely spaced ferromagnetic
nanoparticles within an insulating matrix. Most commonly they are realized in
the form of discontinuous metal-insulator magnetic multi layers (DMIM), (DMIM).
representing a general similarity with the formerly known magnetic
multilayered giant magnetoresistance (GMR) systems (Baibich et a!. al. , 1988).
Compared to the continuous multilayers, the DMIM systems possess the same
advantages (easy preparation, high chemical/electrical stability,
stability. insensitivity
to pinholes) as the cermet films, that is the arrays of magnetic nanogranules
isotropically distributed in the insulating host (Gittleman et al.,
al.. 1972; Mitani
et al. , 1998). However, but they display a much higher sensitivity of MR to
low and moderate magnetic fields than cermets. An example of a DMIM
system is given in Fig. 6. 1, showing a cross-section micrograph of
nanogranular CoFe layers 'in the insulating AlAbz 0 3 matrix.

Figure 6. 1 Cross-section micrograph of granular CaFe layers (dark spots) in Ab


Aiz 0 3 matrix
(brighter background) .

DMIM preparation is based on the fact that if the surface energy of the
metallic component is much higher than that of the insulator, the metal layer
Nanogranular Layered Magnetic Films 183

does not wet the oxide and, below some thickness, it breaks up into
nanoparticles. Besides the co-deposition, typically used for preparation of
cermet films, the alternating deposition of oxide and metal can be employed
for DMIM also. This alternating deposition can be much better controlled than
co-deposition, and in the DMIM case one may tune ideal conditions for each
component separately.
So far diverse DMIM systems have been studied: ColAl Co/Al z0 3 multilayers
multi layers
(Morawe and Zabel, 1995; Naudon et al., 1998; Babonneau et al. , 2000)
and Aiz03/Co/AIz0
Aiz 0 3/Co/ Aiz 0 3 sandwiches (Schelp et al., ai., 1997 and 1999; Desmicht
et al.,
ai., 1998; Fettar et ai., 1998; Briatico et ai., al., 1999; Maurice et al. ,
1999), and the combinations of CoFe/HfOz (Sankar et ai., 1997>, Co/SiO
al., 1997), Co/SiOzz
(Sankar et ai., 1998; Dieny et al., 1998; Schaadt et ai., 1999), Co/ZrOz
(Giacomoni, 1998), Col AIN (Zanghi et al. , 2001), Fe/ AIz03 (Kumar et al. ,
Fe/AIz03
2001), Fe/ZrOz (Auric et ai., 2000) and Fe/CaFz (Mosca et al., 2001).
Typical thickness values fall in the range O. 5 - 2 nm of metal layer and
3 - 4 nm of oxide layer, and sometimes annealing at temperatures between
600C complemented the processes. Most of these systems were
200 and 600'C
prepared by magnetron sputtering, except for (Zanghi et al. , 2001) pulsed
laser ablation and (Mosca et al. , 2001) electron beam evaporation.
The first indication that in Col Al z0 3 multi layers the Co layer gets discontinuous
multilayers
at thickness less than 25 A A was given in 1995 by Morawe and Zabel (morawe (morawe and
Zabel, 1995). It was then found that all the metallic layers almost equally contribute
to the electrical conductivity (even when tco > 25 A), despite the fact that only one
Co>
boundary layer is atomically close to the electrical current leads and no short circuits
between neighboring Co layers were found with high resolution transmission electron
microscopy (HRTEM). It was also concluded that metal-insulator atomic
intermixing might only occur at certain heavily distorted sites within the
sample.
The formation of a single layer of nearly spherical Co nanoparticles in an
insulator matrix was then demonstrated in Col Al z0 3/Co tunnel junctions
prepared by sputtering (Schelp et al., ai., 1997), where such a layer was
inserted inside the Al z0 3 spacer between the Co electrodes. Later, systematic
studies of the Co layer structure as a function of its thickness were performed
using the high-resolution transmission electron microscopy (HRTEM) technique
(Fettar et al., 1998; Briatico et al., 1999; Maurice et al., 1999). The
detailed analysis indicated that the coalescence regime can be divided into
two stages: the first one for the layer thickness t Co < 7 A, where the metal
aggregates are almost spherical and their structure is mostly fcc, and the
second one for 7 A co <20 A, where the aggregates become more and more
< t Co
A<t
elongated and hcp structured. Correspondingly, the relative standard deviation
of sizes reaches a minimum of 25% 25 % at t Co = 7 A. The critical cluster diameter
co =
associated with the referred change of shape and structure falls in the 27 - 30
A range. Finally, the Fourier analysis of TEM micrographs indicated that the
A
cluster arrangement is not random, but that the distance between nearest
184 G. N. Kakazei et al.

neighbors is almost constant at a given nominal thickness. When the thickness


increases, the cluster size scales with that distance so that the width of
alumina spacers between the clusters remains constant (22 A) until the onset
of percolation. Recently it was shown, using the grazing-incidence small-angle
X-ray scattering and HRTEM, that in multilayered ColAI ColAl 2 3 films (prepared
z0 3
under identical conditions) in a certain range of thickness, the vertical
arrangement of clusters from plane to plane is not random but evidences a
topology-induced self-organization (Babonneau et al. , 2000).
Magnetoresistance studies on DMIM started with [CoFe/HfO [CoFe/Hf02 z IN and
[Co/Si0
[Co/SiOzJN2IN (where N-number of bilayers) systems prepared by magnetron
sputtering (Sankar et al.,
aI., 1997, 1998; Dieny et al., aI., 1998), leading to room
temperature MR values, both in current perpendicular to plane (CPP) and current in
plane (CIP) geometries, up to 4% 4 % after optimized annealing. The highest value was
found in the [Co/ Si0
SiOz2J2
Jz sample at CPP geometry, however the authors mentioned a
poor reproducibility of properties for this geometry. Further optimization of layer
thickness and choice of suitable metal/ insulator pairs could make these systems a
metal/insulator
useful supplement for sensor applications (at not too low applied fields) to the high
sensitive tunnel magnetoresistance (TMR) tunnel junctions (Miyazaki and Tezuka,
aI., 1995).
1995; Moodera et al.,
On our part, detailed studies of structural transport, magnetic and time
dependent properties of discontinuous CosoFezo COSOFe20 / Al z0 3
AI 2 3 multi layers were
multilayers
performed recently (Kakazei et aI., al., 1999, 2000, 2001 a, 2001 b; Sousa
et al. , 2000, 2001a,2001b;
2001 a, 200 1b; K leemann et al. , 2001a,2001b;
Kleemann 2001 a, 200 1b; Petracic et al. ,
2002). The CosoCOSO Fezo
Fe20 alloy was selected for the metallic layer because it is
magnetically soft and possesses a high spin polarization (Bardos, 1969), the
important factor for spin-dependent tunneling. AI 20 3 was preferred to other
AIz03
possible insulators since an enhanced MR ( ..... 10% at room temperature) was
recently reported in sputtered Co-AI-O cermet films (Mitani et al., 1998).
Finally, we used ion beam sputtering since it is more flexible for preparation of
nanostructures than common magnetron sputtering, due to the independent
control of beam parameters (ion density and kinetic energy) and deposition
pressure (Kaufman et al. , 1987).

6. 2 Ion Beam Sputtering

The film deposition was made in an automated load-locked Ion Beam


Deposition (IBD) system (Nordiko N3000) equipped with two broad-beam,
RF-excited ion sources (a"(a' 10 em-diameter deposition gun and a 25 cm-
diameter assist gun) in a "z"
"ZOO configuration. Details on the machine can be
found in (Gehanno et al.,
al. , 1999) and (Cardoso, 2001). The IBD chamber has
a base pressure of 4 x 10- ss Torr.
Nanogranular Layered Magnetic Films 185

Figure 6. 2 shows a schematic representation of the deposition/oxidation


chamber. Six water-cooled targets (150 nm x 104 mm in area) are mounted
around a rotary axis facing the deposition source. The desired target material
can be selected by rotating the axis to a pre-determined angle. A shield is
used to prevent cross-contamination between targets. During the film
deposition the water cooled substrate table is at 80, relative to the beam
direction. The substrate holder rotates via a Ferrofluidic seal at a selected
speed between 0 - 30 r/ min, automatically controlled through a de servo drive
r/min,
with encoder feed back for precise positioning. Mounted around the substrate
table is a permanent magnet array that provides a 40 Oe magnetic field with an
al ignment uniformity better than
alignment 2
2 over a 4 area .
rjJ4 inch diameter area.

.Sampling chamber

0d-
Ar 0,

-'-'-'-'-'-'
O{___
-O'-Ar"--_

".
11O
~~~C'~;?:> Xe

._._._._._._.
N

. . . . NOOJ":~:~'-'-'-'-'-'
_._._._._._._.
1

Cryo Turbo
pump
pump pump

Figure 6.2 Schematic of the IBO


ISO system. Pictorial representation of the plasma during
N,1 and N,
oxidation is shown. N N2 are the neutralizers for deposition and assist guns,
respectively. Sample distance to assist gun grid and target are 30 and 20 cm,
respectively.

6.3 Samples Preparation and Characterization

Ab 0 3 multilayers was done in a total automatic


The deposition of the CoFe/ Ab03
process. A max,imum of 8 substrates were mounted on metal platens and
loaded (one platen at a time) directly from the cassette to the table by the
transport arm. The platen was then clamped to the elastomer-coated surface,
ensuring good thermal contact with the water-cooled substrate table. During
186 G. N. Kakazei et al.

the sequential deposition of the layers, the table rotated at 15 r/min with an
80pan angle (beam-sample surface)
surface)..
The Al z0 3 films were prepared from ion beam sputtering from an AizAl z 0 3
ceramic target. The deposition was done with Xe gas (1 sccm Xe) in the
deposition gun, and a beam current of 33 mA. In this case, no extra oxygen
was injected inside the chamber for stoichiometry correction, as in the case of
standard Al z0 3 films deposited in this system. Target properties are described
in Table 6. 1. Both for CoFe and Al z 0 3 film deposition, the Xe ions from the
100 W RF plasma created inside the deposition source were accelerated by
a ,.+
.+ 1450 V voltage applied to the positive grid, and extracted with - 300 V
applied to the negative grid. For the oxide layers, a filamentless neutral izer
neutralizer
(40 mA, 3 sccm Ar) beside the deposition source was used, avoiding target
charging during the dielectric deposition. Normal pressure during deposition
10 - 55 Torr.
was 3 x 10-

Table 6. 1 Target properties.


Material Target Thickness( mm)
ThicknessCmm) PurityC %)
Purity( Deposition rate( A/ s)
rateCA/s)

go Fe2o
Coao
CoFe 2 99.95
9995 0.38
038
PureTech

Ab03
Ab03 6.4 99.9
999 o 120
0.120 (80
C80 pan)
SINEX

Table 6.2
6. 2 Properties of 600 - 3000 A thick AI
Ab2 0 3 fi Ims.
3 films.

Deposition gun: 33 mA, 1sccm


lsccm Xe,
V+ =
== 1450 V, V_ =
== - 300 V
Deposition conditions
Deposition neutralizer ON (70 mA, 3 sccm Ar)
Assist gun: 10 sccm Ar-15 %O
% O2 , no plasma

Breakdown voltage 2 3. 6 x 106 V/ cm


2.. 3 - 3.
step coverage 58 % lateral step coverage
refractive index 168
68
deposition rate 7.2 A/min
7.2A/min
thickness loss in 10%
10 % HF solution 700 - 800 A/min

Table 6. 2 summarizes the properties of 600 - 3000 A A thick Aiz 0 3 films


deposited by IBD.
IBO. The standard process uses the assist gun gas inlet as an
extra source for oxygen (3 sccm),
sccm) , for a stoichiometry correction.
The composition of oxide films was analyzed by Rutherford backscattering
analysis (ITN, Sacavem) and the metall ic impurity concentration (Mo, W,
metallic
Mn, Ir) was found to be less than 400 ppm. This contamination is
characteristic for the ion beam systems, and comes from the sputtered
Nanogranular Layered Magnetic Films 187

materials of then grids, from the shields surrounding the targets, or cross
contamination from the neighboring targets.
The refractive index was measured on oxide films deposited on Si
substrates, with a Rudolph Research/Auto EL ellipsometer model IV-NIR 3,
using a tungsten halogen lamp, and a wavelength i\ = 632. 8 nm.
All the CoFe/AI
CoFe/ Ab20 3 films were deposited on glass substrates, with a Xe
ion beam acting alternatively on two separate targets. The"The'insulating
insulating layers
were directly sputtered from an Ab 0 3 target, with a deposition rate of
O. 12 AIA/ s. COao Fe20 (the composition confirmed by the Rutherford
backscattering data) was sputtered from a mosaic target (pieces of Fe on a
A/ s. The nominal thickness t of the
Co plate) with a deposition rate of O. 32 Als.
CO aQ Fe20 layers was varied by steps of 1 A from 10 to 14 A, by steps of 2 A
from 14 to 20 A and finally t = 25 A. The thickness of the Ab 0 3 layers was
fixed at 30 A. For this whole thickness range, we systematically prepared
films with a fixed number of bilayers n = 10. In addition, for t = 13 A we also
prepared samples with n = 1, 4 and 7. Layer thickness was determined from
known target cal ibrations and subsequently confirmed by low angle X-ray
calibrations
diffraction to be within less than 5%5 % of expected values. We always used
Ab03 as bottom and top layers, so the number of Ab 0 3 layers was n + 1' 1.

6. 4
6.4 Structural Properties

The structural properties of these films were studied with low angle X-ray
diffraction, exhibiting the spectra characteristic for multi layers . The interface
quality improves with increasing CoaoFe2o layer thickness (Fig. 6.3) 6. 3) as
revealed by the better defined superlattice Bragg peaks, distinguishable up to
5th order for t == 20 A. They are indicative of improved lattice coherence, as
CoaoFe2o layers get thicker and eventually continuous. Also, the more
pronounced Keissing fringes indicate improvement of the structural quality and
surfaces. As seen below from the conductivity data, at
flatness of the external surfaces,
tt~
~ 18 A the CoaoFe2o
COao Fe20 layers are continuous, while below this thickness they
become discontinuous giving rise to spin dependent tunnel transport between
metallic clusters. However, even CoaoFe2o layers as thin as 12 A still give rise
to three low angle Bragg diffraction peaks (Kakazei et al. , 2001 a) .
The size and dispersion of these metallic clusters were studied in high
resolution transmission electron microscopy (HRTEM) experiments, performed
on a Philips CM30 system with point resolution of 1. 9 A at 300 kV. To enable a
reliable estimation of the granule size distribution, a sample with a single
CoaoFe2o
Co ao Fe2o layer within the Ab 0 3 matrix was prepared for in-plane view on' a
large area. The micrograph in Fig. 6. 4a is a plane view HRTEM image
6.4a
=
obtained on the sample with a single t 13 A CoFe granular layer (in order to
188 G. N. Kakazei et al.

~l ----'------------'-----'--------'---

1:i~4l~
4l ~. 1=16A

,~~--,!-----.
1=16A
6 3

...J
2 -:-.....,.-....-~
L -_ _--'- ...L-_ _---' _
...J

:l o
LI

2
, ' - -_ _--',,.---_ _

4
2en
6
I=_1--'-2_A_ _--'1
8
,
10

Figure 6.3 Small angle X-ray diffractometry on CoFe/ Ab


AIz03
0 3 films with continuous (t =
==
20 A) or discontinuous (t = 16 and 12 A) CoFe layers.
(t==

60,---.-----.---,---,---,-----,
60

50
Iii
1&

*::l

~o
40
-Iii:

30 _I""
'-
(1)
.D
E 20 I"
::l
Z i\,
I"
10

o 5
iii .finlrrhM
10 IS
15 20
f1--f?l
25
Diameter(nm)
Diameter (om)
(a) (b)

Figure 6.4 In-plane view (a) and granule size distribution histogram (b) for a t =
== 13 A
A
Ab2 0 3 system.
single-layer CoFe/ AI

avoid the superposition of images from different layers). It clearly evidences


the presence of numerous small precipitates embedded in an amorphous
matrix. Image analyses were carried out on several HRTEM micrographs from
various regions in this sample to get the size distribution histograms of these
clusters. As seen in Fig. 6. 4b the size distribution has a typical log-normal
6.4b
Nanogranular Layered Magnetic Films 189

shape with a maximum centered around 30 A and it extends up to 300 A.

6. 5 Magnetic Properties

6.5.1 Nuclear Magnetic Resonance


Important information on the short range atomic order within the magnetic
component of the obtained DMIM systems is provided by the nuclear magnetic
resonance (NMR) spectra on Co nuclei, which, in this case, demonstrate new
specific features compared to the well studied cases of bulk or continuous thin
film CoFe alloys.
NMR spectra have been recorded in a zero external dc field at 1. 5 K using
a phase-coherent swept frequency spin-echo spectrometer. Fig. 6. 5a - e
show the NMR spectra recorded in a series of [Co eo Fe20 ( t) / Ab 0 3 (3nm) ] 10
[Coso

2 I -II A
I-IIA 1.2
1

,
2-12A

0.4 C
0.8
st -- FeCo 16 16~~
~ 0.4 ---n- FeCo 20
Pure fcc Co -{]-
~ FeCo 18A
~FeCoI8A
A
0
150 200 250 300 350
350 190 200 210 220 230
(a)
(d)
2 1-13 A
1-13A
~ 0.8
,-..

,
2-14A
:i~
~
Q;
'E 0.4
E
.~
.;;:;
B
~
c:
OG.-_ _-'-__
0 L......_ _--'-_ _
C

--'-_---.:::L=~_
.L..-_..:::c~~_
~
00 ~
...~
...... . .. ...
.=
.: 150 200 250 300 350

~
_~D+'"
o
~
"@ (b) cfXJo ~ o~...
c: 0 00 0 000D;-/lt'
ocJflJOo
o0
~
.0.. .... 0 . 0
VJ
C/l ... 0 ..,. 0 0 0 0
.... 00
1-16A
l-16A 00 0
0.8
2-18A ....... t
....... A
6oa::J:P
-0- FeCo 11 A
FeCo 12A
0 0
0
0
3-20A --0-
:or.
; . Pure bee Co
-+- FeCo 13 ~ 0
0
--T- FeCo 14 A

200 250 300 350 190 200 210 220 230


Frequency (MHz)
(c) (e)

Figure 6.
6.55 NMR spectra in CoFe/ Ab
Aiz 0 3 multi layers at different thickness of
ferromagnetic layers.
190 G.N.
G. N. Kakazei et a!.
al.

films (a) - (c) and parts of these spectra (d) - (e) in the central 190 to 235
MHz range. These spectra can be generally analyzed within three
characteristic patterns, associated with three qual itatively different types of
magnetic layer structures, each of them related to a specific thickness range
as t is varied from 8 to 20 A: (1) t< 13 A(Fig. 6. 5a); (2) 13 A< t< 16 A
(Fig.6.5b); (3) t>16A(Fig.6.5c). The spectra are similar for samples
within the same region and change rather abruptly when passing to the next
one. The physical interpretation of these types was made in terms of specific
contributions from Co atoms inside the granules or at their interfaces with the
amorphous AI 2 0 3 matrix, as described below.
Ab03
The simplest NMR spectrum was obtained in the thick, t = 20 A, =
practically continuous film corresponding to region 3 (Fig. 6. 5c, curve 3 and
Fig. 6. 5d, open squares). Curve 3 displays the narrowest resonance line
among all the spectra in Fig. 6. 5a - c, with the peak at 220 MHz. One also
observes a large tail towards high frequencies and a low plateau in the 160 to
180 MHz range. This broad spectrum and its high frequency tail are typical of
CoFe bulk disordered alloys with 20 % Fe content, both fcc and bcc (Wojcik
et ai.,
al., 1997). Actually a mixture of the two phases is expected from the bulk
phase diagram.
One could compare the NMR signal in the low frequency range to that
earl ier observed in bulk CoFe ordered compounds (8 2 phase) (Jay et al. ,
earlier
1996), where it was attributed to misplaced Co atoms lying in the Fe
sublattice of the 8 2 structure (antisites). However our NMR spectrum does not
lines typical of such a 8 2 phase (Jay et ai.,
show the multiple narrow Iines aI., 1996).
Therefore the low frequency plateau must be attributed to Co atoms located at
the metal-insulator interfaces, where the hyperfine field is lowered due to the
presence of non-magnetic neighbors. As seen below, this is confirmed by the
evolution of the spectral shape in the thinner samples.
The next thinner film (with t = 18 A) exhibits a similar spectrum,
although sl ightly broader. The signal intensity in the range from 160 to
180 MHz is the lowest in all samples, indicating minimum interfacial
roughness. This is confirmed by transmission electron microscopy, which also
shows that the CoFe layers are continuous in this thickness range. The
spectrum for t = 16 A is generally similar to the previous ones in its main part,
but already shows an increase of the signal in the interfacial-related 160 to
180 MHz frequency range. This indicates that CoFe layers start to break up at
such thickness, thus increasing the surface area. Therefore the continuous-
discontinuous film transition can be practically fixed at t = 18 A, the so-called
structural percolation thickness.
At the opposite end of the CoFe layer thickness in our series (region 1),
the NMR spectra behave Iike like those shown in Fig. 6. 5a, e (open circles and
squares), for the films with t = 11 Accurve
A (curve 1) and 12 A A(curve
(curve 2). The signal
intensity in the 160 to 180 MHz range dominates in these spectra, indicating a
very high surface/bulk ratio: the samples consist now of granular CoFe layers
Nanogranular Layered Magnetic Films 191

with a large interface area. The large interface area is also consistent with the
weak feature appearing in the spectra at ....... 330 MHz, a frequency which is too
-330
high for any CoFe alloy (even diluted Co in Fe). Most probably, the latter
signal comes from the enhanced Co magnetic moments near the interface due
to a transition from band to localized d-electrons at the formation of CoO. This
resonance signal was absent in region 3 (Figure 6. 2c) where the relative
granule surface area is much less than in regions 1 and 2. A similar signal at
enhanced frequencies was observed in Co nanoclusters embedded into a SiOz
matrix (Thomson
<Thomson et al. , 1997).
Besides the above-discussed characteristics mainly related to the
interface Co atoms and granular/discontinuous features, the spectra also
contain a "common" bulk contribution from Co atoms inside continuous layers
or inside granules. This "bulk" contribution is associated with resonance peaks
of about equal intensity in the vicinity of 200 and 225 MHz and a hump at higher
frequencies (-250 ( ....... 250 MHz). Such a set of three lines corresponds fairly well to
the spectra observed in thin bcc CoxFe ,_x films at the relevant composition
COxFel_x
(Wojcik et ai., al., 1997>. 1997). They correspond to Co environments with dominant
even numbers of Fe neighbors, 0, 2 and 4 respectively; a kind of short-range
order that is not observed in bulk disordered and ordered (B z ) CoFe alloys.
Therefore in the FeCo/ Al z0 3 multi layers we rather observe a mixture of
unknown bcc phases as in thin CoFe films.
The NMR spectra in the intermediate region 2 (Fig. 6.5b, 6. 5b, d, solid
triangles and diamonds) display an abrupt decrease of intensity at the
"interface" ranges, 160 to 180 MHz, and and""'"
- 330 MHz, in comparison with the
spectra in region 1, and a corresponding increase in the "bulk" range of 220 to
255 MHz. This, in addition to the above remarks, excludes the Bz2 "" antisite"
origin of the low frequency signal. Indeed, in the case of off-stoichiometric Bz
phase the intensity from this region should evolve like the intensity from the
region with frequency f> 250 MHz, which is absolutely not the case. This
confirms that the signal below 180 MHz arises purely from the interfaces.
Moreover, it decreases similarly to the high frequency peak C, due to oxide
formation at the interface. It should be noted that the decrease rate of the
interfacial regions (metallic f< 180 MHz; oxide f>300 MHz) does not obey
the 1/ t law expected for a continuous film: the surface/volume ratio evolves
much more quiekly and discontinuously.
More structural information can be obtained from finer features in the NMR
spectra. Thus, in regions 1 and 2 one observes a signal at 217 - 218 MHz
corresponding to fcc Co (B in Fig. 6. 5d). This peak starts being resolved in
the film with t = 12 A (open squares) and is shifted towards high frequencies in
the film with t = = 14 A (solid triangles), then in region 3 it is no longer
resolved. The fcc structure in the film of 13 A was inferred from HRTEM data
(Fig. 6.6),
6. 6), giving a lattice parameter of 3.4 3. 4 A. Note that in granules the
lattice parameter can be different from its bulk value. The lattice parameter of
- 3. 5 A and the spectra of the samples of 12 and 13 A are similar
fcc Co is .......
192 G. N. Kakazei et al.

within the 212 - 230 MHz frequency range. An absence of the bcc phase in the
13 A film, as seen from HRTEM data, could be due to the limited investigation
of a small part of the sample only. In fact granules have a broad size
distribution resulting from different local formation conditions, hence a complex
film structure is expected.

6. 6 A cross-section HRTEM micrograph of t = 13 A DMI film and the X-ray


Figure 6.6
diffraction picture from a single CaFe
CoFe granule (dash-rounded) permitting us to identify
an fcc structure with the lattice parameter""" 3. 4 A.

At last, it can be noticed that globally the spectra" gravity center" shifts
to high frequencies when the thickness increases. This is partly due to the
decrease of the relative contribution from the interface. On the other hand, the
variation of the relative intensities of the main peaks at 200 and 225 MHz
shows that the environment with two Fe neighbors (225 MHz) is favored at the
expense of that with no Fe neighbors (200 MHz). This suggests that the
metallic CoFe alloy gets richer in Fe as the thickness increases. In fact, iron
oxidizes more easily than cobalt hence the metallic layers contain less iror.1
near the interfaces because of the chemical reaction between Fe and Al z 0 3 .
This is the same reaction that produces the absorption range C (f ....... 330 MHz)
(f,.."
but it is more important for Fe than for Co.
In summary, this NMR study shows that the evolution of the structure of
Nanogranular Layered Magnetic Films 193

the CoFe layers takes place in three main stages: at t < 13 A the films are
granular; at 13 < t < 16 A they present a mixture of granules and flat local
regions; at t> 16 A the films are continuous. This evolution scheme agrees
with the data on global magnetic order in the same film series, obtained from
FMR and magnetization measurements (see below).
below) .

6.5.2 Room Temperature Static Magnetization


Since the MR properties of DMIM systems, most important for their practical
application, are essentially determined by the magnetic correlations between
neighbor granules, a detailed study was done on the magnetic states in our
films at different values of the thickness t of magnetic layers.
To obtain a first insight, vibrating sample magnetometer (VSM)
measurements of the room temperature static magnetization M (H) were
performed on the whole range of films fi Ims (Kakazei et al.,
aI., 2001 a), the most
characteristic curves being shown in Fig. 6.7. For increasing t one notices a
t~
clear crossover from superparamagnetic to ferromagnetic-like behavior at t"":;;:
13 A where the remanence M,em M rem and coercivity first appear. These parameters

I
o I .:1
..:

_:~~--/:~----
_:13 %~'''----
0.0002 A

o __
.......... II
--{).0002 .
-0.0002 .....:......
.:.' I
)

-20 -10 o 10 20
S
S'
E o
I
~'I'.::---'.: I
....... ..
~ 0.0002

"c:
<l.l
<!)
E
o o0
_:I4A_! _
E
.g<l.l --{).0002
-0.0002
...............:.,.,.,J'::.,I,~I
<!)
c:
"~
:2
::2 -20 -10 o 10 20
I
0.0002 t=18..\
t=18A I
,...,...-r'
~-.-.-.-.-.-.-.-.-.-.-..-.-.

-0.0002
o
o ------+t-i--------
+-lJ
-{),0002 .....
_
I
~
-.-.-.-..-.-.-.-.-... -~
I
I

-100 -80 -60


--60 -40 -20 0 20 40 60 80 100
H(Oe)
H(Gel

Figure 6.
6.77 AI,20 3
VSM magnetization curves at room temperature for [COso Fe20 ( t) / AI 3 (30 A) ]JlO10
films with different values of nominal thickness t of CaFe layer.
194 G. N. Kakazei et al.

exhibit a critical behavior close to t


t' , e.g.,
e. g. , M rem ( t)o::::( t
Mrem(Ooc( t-- tt** ))1'2(Fig.6.
1'2 (Fig. 6. 16)
and magnetization changes drastically in a very narrow thickness range, from
almost purely Langevin behavior (t = = 13 A ) to a clear ferromagnetic jump
under a few Oe applied field (t = = 14 A). As will be shown below, a singular
behavior in the same region t - t * is also exhibited by other physical
characteristics dependent on the long-range magnetic order, as the
ferromagnetic resonance ( FMR) linewidth or the low field sensitivity of
magnetoresistance. On the other hand, the electrical resistivity data (see
below) show that nothing dramatic happens in the nanogranular structure in this
very narrow thickness range where Coso Fe20 layers remains still discontinuous,
as confirmed also by direct HRTEM observation for t = 13 A (Fig. 6.4). 6. 4) .
Comparison between these two types of data gives us an unambiguous
indication of cooperative magnetic behavior, appearing therefore among
disconnected ferromagnetic granules in the system when all other mechanisms
coupling,
of magnetic coupl ing, except long-range magnetostatic, are ruled out. The
corresponding scenarios were broadly discussed in literature, starting from the
classical paper by Luttinger and Tisza (1946, 1947), who stated that the
ferromagnetic ground state is impossible for a cubic or square lattice of point-
like magnetic moments with dipolar coupling between them. However, some
recent theoretical analyses (Prakash and Henley, 1990; Bouchaud and Zerah,
1993) and experimental observations (Takzei et al., aI., 1999; Evoy et al. ,
2000) showed the importance of various additional factors (such as granule
finite size and anisotropy, positional asymmetry or disorder, etc.) that can
bring a dipolar coupled system to exhibit long-range ferromagnetism, even in
the absence of direct exchange interactions. In the particular case of
disordered granular array, such type of ordering is associated with the concept
of "magnetic percolation" (Sankar et al. , 1997; Pogorelov et al. , 1999b), in
contrast with the common structural percolation, responsible for the onset of
metallic electric conductivity.
Further support to such dipolar ferromagnetism in our system can be
directly extracted from the magnetization data. If one tries to explain -- - 25 %
saturation at -2
-- 2 Oe, seen in Fig. 6.7,
6. 7, by a Langevin function with enhanced
superparamagnetic moments, this should require coalescence of -- - 10 3 typical
granules. Such abrupt coalescence would be incompatible with the resistivity
data presented below. Hence the most evident mechanism promoting
ferromagnetic order is the dipolar interaction between granules. This process
will be of our principal interest in the following section.

6.5.3 de and ae SQUID Measurements


To obtain a deeper understanding of the possible magnetic phases in our DMIM
films, we extended the above study to [Coso Fe20 ( t) / Aiz
Ab 0 3 (3 nm) ] 10 samples
with different values of t, using ac and dc SQUID measurements at
Nanogranular Layered Magnetic Films 195

temperatures 10<T<300 K, ac frequencies O. 01 Hz <f<10 3 Hz, and an ac


field amplitude h = 4 Oe in a virtually zero external field (Sousa et al. , 2001 a;
Kleemann et al. , 2001 b) .
Figure 6. 8 shows the data obtained for four different DMIM at several
frequencies O. 1::(; 1~f~100f::(; 100 Hz. For the t= t = 10 A
Asample (Fig.6.
(Fig. 6. 8a) the real and
imaginary parts of magnetic susceptibility X, X' (f, T) and X" (f ,,T), T), are
similar to the data observed previously on frozen FeC ferrofluids with a low
density of ferromagnetic particles (Jonsson et al., 1998). A sizeable
frequency dispersion in the range 40::(; 40~ T::(;80
T~80 K can be associated to the onset
of superspin glass (SSG) phase, while a non-dispersive Curie-weiss (CW)-
n
type decay of X' (f, T) with an extrapolated ferromagnetic (FM) Curie weiss
temperature : : :;:
e """" 58 K is encountered at T > 80 K (see 1/ X' x' curves for
= O. 1 Hz and the best-fitted CW plot within 200::(;
f= 200~ T::(;300
T~300 K in Fig. 6. 8a). At
higher nominal thickness, tt;;? ~ 11 A, a dispersionless background appears at
higher temperatures, in addition to the polydispersive response curves of the
glassy subsystem at low T Fig. 6. 8b - d). Its high- T tails can be described by
T((Fig.
a CW law with FM mean-field Curie temperatures e( t) """" 114, 165 and 270 K
t):::::;::
for t = 11, 12 and 13 A, respectively (see 1/ X curves for f = O. 1 Hz and CW
plots best-fitted above 240 K in Figs. 6. 8b - c). These estimates agree well
with the appearance of ferromagnetic behavior at room temperature in the 14 A
sample (Fig. 6.7). 6. 7). Hence we interpret the background curves as being due to
the prevalence of FM "exchange" (due to stray dipolar fields from non-
spherical granules, see below) between" magnetically percolated" granules
over pure dipolar coupling coupl ing (Luttinger and Tisza, 1946). Their aggregate (the
infinite cluster) behaves Iike like a FM with finite in-plane anisotropy, which
causes the susceptibility to decrease upon cooling below the FM ordering
temperature, T c < e. Notably, the dispersionless part of X' ( f, T) is not
linked through a usual scaling relation (Ogielski, 1985) with the energy losses
estimated from X" (f , n. T) .
To specify the nature of this FM state, which can be also called
superferromagnetic (SFM) as resulting from FM ordering of SPM moments, we
have measured the de dc magnetization M - 45~ H ~
( H) in the low field range -45::(;H::(;
M(H)
45 Oe at different temperatures obtained after zero-field cool ing (ZFC) from
300 K. The field step was 1 Oe for t = 13 A and t in the interval from 220 to
260 K, but also an enhanced resolution (0.05 (0. 05 Oe) was used at t descending
from 300 to 80 K (t = 12 A), and to 120 K (t = 13 A), respectively. Each of
the curves (partially shown in Fig. 6. 9) was obtained after zero-field cooling
from 300 K. Langevin-type SPM magnetization curves are observed at high t,
but upon cooling to below a certain Curie temperature T c they develop finite
jumps t:.M H:::::;::O. The particular T c values determined were""""
J::,.M at H""""O. were :::::;::150
150 K (for
= 12 A) and :::::;::240
tt= """"240 K (for t = = 13 A). Upon further cooling t:.M( J::,.M (T)T) grows
accordingly to the stabilization of SFM order and eventually defines the
growing M rem rem ( t) .
196 G. N. Kakazei et al.

~
~

7
0
,...,
<--' 4
~
-~

0 t=JIn:mtnmtlmrnl:rorrl=al 0
100 200 300
Temperature (K)
(a)
8

~
~

7
0
,...,
<--' 4
~
~

0
100 200
Temperature (K)
(b)

7
0
,...,
<--'
4 ..
0
~
-~ <--'
-~ ~
~

0
100 200
Temperature (K)
(c)

Tg
8
~

'I'
0
--0.1 Hz .
~

0
<--' --1.0Hz <--'
-~ 4
--10Hz ~
~ --100Hz

o0 100 200 300


Temperature (K)
(d)

Figure6.8 X'Cf, T) and X"Cf, XHCf, T) vs. T of CoFe/Ab03 DMIMs with t= 10Ca), 10(a),
11(b), 12Cc) and 13 A (d)
11Cb), 12(c) Cd) measured at frequencies 0.1 < ~ f<f~ 100 Hz. Note the
( f = O. 1 Hz, T)
H

magnification factor (5x)


C5x) applicable to all X"
X data. Inverse curves X' -1 (f
CCW) lines define the mean-field transition temperatures, e,
with best-fitted Curie-Weiss (CW) B,
as marked by arrowheads together with glassy temperature Tg, T g' T, and T c c)
CCc) and (d)
Cd)
only), so that Tg<T,<Tc<e.
Tg<T,<Tc<B.
Nanogranular Layered Magnetic Films 197

Obviously, the SFM system behaves like a very soft ferromagnet. Just
below the respective T c, e, it is demagnetized in zero field and may be switched
into its spontaneous values l:.M flM = = M s (spontaneous magnetization) by
applying external fields in the order 1 Oe. The quantity (flM(l:.M// flH) max vs. T
l:.H)max
taken from the M(H) curves measures SPM susceptibility at T> T ec . At T<
l:.H within less than 1 Oe) to the jump l:.M,
T c it is proportional (at flH flM, which is
a measure of Ms.

=1
S'l
E
Q)
~

'"Q)
E
40

'h
l, Ol---~---c:
0 I---~--::; ilIf""F-------I '7 1------",,-------t#if-------iO
o
x x
~-l I -""n'"""..-
~
-40
-2

o 2 4 -4 -2 0 2 4
J1PoH (mT)
oH(mT) J1 oH(mT)
poH(mT)
(a) (b)

Figure 6.9 Low-field magnetization curves M vs. fJo H obtained on the DMIM with t = 12
Ca) and 13 A (b)
(a) Cb) after ZFC from 300 K to (a)
Ca) 300, 260, 220, 180, 140, 120, 100 and
80 K and (b)
Cb) 300, 260, 220, 180, 140, 120 K.
K, respectively.

At a certain "freezing" temperature T f < T c.


T,<T e , the M (H) curve turns to be
M(H)
hysteretic as evidenced by the finite values of M Mrem and coercive field He
(Fig. 6. 9). For instance.
instance, we defined freezing temperature T, T f ~ 120 K for
=
t = 12 A and ~ 190 K at 13 A. Obviously.
Obviously, below this temperature the free
energy barrier due to the weak intraplanar anisotropy is no longer overcome by
thermal activation, and both M rem and coercive field He increase upon further
cooling. Since the T T,f values roughly coincide with the onset of low- T
dispersion in the ac susceptibility data (Figs. 6. 8c - d), the observed
hysteresis may be related to the metastability of the re-entrant superspin glass
(RSSG) phase. Indeed, there is an observable decrease of the hysteresis
with waiting time t w between the data points, e. g., He goes from O. 85 to
= 13 A
O. 65 Oe in the t =
0.65 A sample at T == 150 K when t ww increases from 90 to 900 s.
This suggests some link between the hysteresis and lossy dynamics of the SSG
component, also seen in its frequency dispersion (Fig. 6.8).
Finally, under yet deeper cooling to some "glassy" temperature T g < T T"f
the hysteresis reaches saturation while the energy losses X" (f, ( f, T) reach a
maximum. We associate this with the SSG transition in the subsystem of
magnetically non-percolated SPM moments. Thus, the emerging SSG phase is
spatially separated from the SFM phase, allowing for coexistence of the two
phases in such a "re-entrant" state (unlike transformation of FM into SG in
usual re-entrant transitions (Srivastava et al. , 1994.
1994)).
To treat the temperature dependence of ac susceptibility, we modify the
198 G. N. Kakazei et al.

known Neel-Brown approach (Brown,


CBrown, 1963), supposing each granule moment
/.1 is subjected to a certain random uniaxial anisotropy constnt A and random
J.I
local field H, all lying in the DMIM plane, with the energy:
E(e) = AJ.l2
E(8) in28
A/.12 ssin e -- J.lHcosC8
/.1Hcos(e - e
8H ) (6.1)
C6.1)
e
where 8 and e 8H are respectively the angles that J.I /.1 and H make with the
anisotropy axis. Depending on the ratio between Hand H A = 2AJ.I' 2A/.1' the function
E(e) can have one (for H>H A ) or two (for H<H A ) minima 0~81<82~n,
E(8) 0~e,<e2~n,
iwt
h",w Ct)
and in the latter case the ac field h he;""
( t) = he ,w = 2nf, can produce non-
equilibrium occupation of the minima and give rise to energy losses. The
kinetic equation for the occupation number n 1I = 1 - n2 = n of the metastable
minimum is:
C6.2)
(6.2)
where the escape transition rates Y 1.2 = Yo expC exp( - f3h/3h 1.2 ) , inverse temperature
/3 = (Ck B T) - 1, and the "attempt" frequency Yo '" 1
f3 08 S - 1. But there are also
10
some small contributions, due to ac field h"" h w ' into instantaneous values of
energy barriers hl. h1.22 for escape from metastable and stable minima. This can
be presented as h h\~~ + KKI.2
hl.1.22 = hj~~ 1.2 Reh w'
w ' with some constants of field
sensitivity of energy barriers K 1.2' 1.2' bringing some time dependence into y 1.2 .

The general solutIon


solution to Eq. (6.C6. 2) is:
I, t ,-

nCt) = exp
net) ex p [ - frcndt']fex p[
frU')dt']fexp [ -- frcndt]Y2(t')dt
fnt")dt]Y2()dt (6.3)
C6.3)

rate r= Yl + Y2' and, after a waiting time twr-
with total transition rate,= tw,-I,' , it
attains a steady state regime n Ct) ( t) -+- n s-ss C
- ns- ( t), which is the most important in

practice. It is the difference between ns- n frSs and the equilibrium value no = {1 +
exp[ f3~J}
/3.t.]} -1 (where
Cwhere .t.
..1 is energy difference between stable and metastable
..1 = h~O)
states, .t. hiO
h\O that defines a contribution to complex ac
susceptibility Xac:
(6.4)
C6.4)

including the average over time and orientations of random fields and
anisotropies. At high enough T, such that , w, the steady state is simply
r
ns-sC t)~ { 1 + exp[/3.t.(
ns-s(t)~{ exp[f3~ Ct)]}-1 ~ Ct) = h 2
t) J} -1 where .t.( 2 C
(t) Ct) adiabatically
t) - hI (t)
follows the variation of the energy levels. Then we obtain:

C6.5)
(6.5)
and the self-consistency relation R = JXh w' where J is the above used FM
"exchange" constant, leads to the CW law with e", J / k B ' " T c .
Equations (6.3) (6.4) still apply for a more complicated non-
C6.3) and C6.4)
C, '" w) taking place at T'" T g . Here we only note the
adiabatic case (r
3
estimate for energy barriers'" k BB T
Tgln Yo// w) gives'" k BB x
gin (CYo X 10 K for our
system. This is too high to be accounted for by the typical interaction energy
Nanogranular Layered Magnetic Films 199

of two granules J./ J.12 1


/0a 33 ' " k 8B X 102 2
K (a is mean distance between neighbor
granule centers), and gives a hint that effective SPM moments (superspins)
should involve up to '" 10 granules.
Finally, the phase diagram of magnetic states in the t- T variables
(Fig.6. 10) can be summarized as follows. The SPM-SFM transition (line 1 in
Fig.6. 10) occurs when the dipolar interaction energy, which is continuously
frustrated for point dipoles (Luttinger and Tisza, 1946) or spherical granules,
obtains a growing contribution from stray fields produced by non-spherical
particles. The latter appear due to coalescence at growing size d = (6ta (6t0 2
2
1/
IT) 1/3 of spherical granules, nucleated at random centers with mean distance

a.
o. The relative number of coalesced granules rapidly increases with nominal
thickness t as p ( t) = 1- = 1 - exp[ - ( t /0) 1a) 1/3 J, hence their structural percolation
occurs at pCts)~O.
p ( t s) "'::::0. 5 and tts~a/3,
s "':::: 0 /3, while the magnetic percolation begins at
only slightly lower p( tm)"'::::O. tm)~O. 45 but at notably thinner ttm~a/5. m ",::::0/5. The effective
Curie temperature rapidly grows as T Cc (( t) oc 2 2
0::; t ( t - t~), eventually reaching
( t s) ' " T C.bulk ' " 103 K when the infinite SFM cluster turns into usual FM.
T cc (t

l~r1
1000
I
900
1 TCbulk
I
I
I
~ 400
I
~ I
'"~ I
e5 300 SPM
:::l
I
SFM I FM
E"'" 200
'" I
E I
~ I
100 2 \
0'-----_
0------_
\: I

~...[]
~--o
I
~I
"I
0 - - -SSG RSSG
0
0.6 0.9 1.2 1.5 1.8 2.1
Nominal thickness (nm)

Figure 6. 10 Magnetic phase diagram of Coso Fe20 ( t) / Ab AI 20 3 (3 nm) DMIM in the


variables t and T: (0) SPM-SFM transition by Xdc' Xde' XacXae and MR data; (0) (D) onset of
ae. Theoretical curves: (1) T e cc 0:; t 2
4/3 - t';(3);
t'/,3); (2) T gO:; 2
SSG state.
state, by X ac' 2 (( t ,3 9 cc t ;

(3) TgO:;(
TgCC( tt- 001''''''1300 K (as for bulk CoSOFe20)'
t,)PI2; T e . bUlk",=,1300
- t,)p/2; COSO Fe20 ).

A rough estimate of the temperature T 9 ( t) for t< t mis m is given by the mean

fluctuation of the dipolar field'" J/J.12 1a 3 , resulting in its t-dependence T 9g (t) '"
/0
(6M/lT)2 ot
(6M/IT)2 at 22 (SPM-SSG, line 2). This increase of T g (t) is limited at t> t m m
and drops to zero oc 0::; (t
( ts - t)Vv at tt-- t s (SFM-RSSG, line 3), due to the
t)
vanishing probability to find SPM clusters in an almost FM medium, and this
e. g. , by the percolation critical index v=
can be described, e.g., v = 5/36 (Stauffer,
1979).
1979) .
The SFM ordering can be responsible for the above referred observations
of magnetic remanence and coercivity in different granular magnetic systems,
both disordered (Bouchaud and Zerah, 1993) and ordered (Takzei et aI., al. ,
1999) structurally. For a disordered granular structure, the magnetic state at
200 G. N. Kakazei et al.

t>t'
t> t * should present in general a mixture of two coexisting phases: a dipolar
coupled ferromagnetic phase (SFM) having the relative fraction f F = M rem (t)
( t)//
( tM coFe ) where M CoFe is the magnetization of continuous CoFe film, and a (1 -
f F) relative fraction of residual SPM phase. The latter is the most probable
source of the weak response seen at H H >4 Oe in Fig. 6. 7b.

6,5.4
6* 5* 4 Ferromagnetic Resonance
In order to obtain dynamical information on the magnetic state of these
systems, the ferromagnetic resonance (FMR) has been studied in multilayered
Coso Fe20 ( t )/ AI 2 0 3 (30 A) samples with t varied from 8 to 25 A(Kakazei
et al. , 2001b).
2001 b). X-band FMR spectra have been recorded at liquid nitrogen
and room temperatures using a standard EPR spectrometer with a two-axes
sample holder. For all t values, the FMR field H,H r showed no azimuthal angular
dependence. Its polar angular dependence was treated using standard Kittel
formulas and the inferred effective anisotropy field H eff as a function of t is
Hell
shown in Fig. 6. 11. This function clearly displays three different regions,
separated by the same characteristic values t 1,2 that were reported in previous
magnetotransport studies on this system (Kakazei et aI.,al. , 1999, 2001a).
2001 a) .
,
,
., 25 ,
,
"
0
~ 20
620 /
2'
, '\,II
~ '"
::t:
v
\
"0
"v 15 ~
~ ',/\-\
\ (I-I

"2 10
t;:
>-. "1
e,'
I-i
r-2
I

-,".
c.. _,' 1,2
e,".
. ..
2
.~'0'" I' .....
.....

'"
'""">
oj
<l)
5
I'
I ""
,
,1 '' I
I e- ....
..
...
"n
'B 0 , I I

~ I, 1I
/
w SP
SP'I M FM FM 1M,
I M I SP
/
-5 I I

0 5 10 15 I 20
15' 25 0.3 0.6 0.9 1.2
(A.)
CoFe layer thickness (A)
CaFe Inverse thickness (l/nm)
(1/nm)
(a) (b)

Figure 6. 11 Effective anisotropy field H eff ( t) at room temperature reveals three


different regimes with respect to nominal CoFe layer thickness t. (a) Linear
t-dependence for SP phase at t<t" t<t l , and (b) linear t-'-dependence
t-I-dependence for FM phase at
t-'<ti
t-'<t2". l
The interval t 1l <t<t 2 pertains to the mixed (M=SP+FM) phase.

In the region t 2 <t, a linear dependence of H eff on inverse thickness t- 1 is


observed (Fig.6.
(Fig. 6. 11b), common to continuous ferromagnetic multi layers under
the mutual effect of demagnetizing field 41fM,
4rrM, volume anisotropy constant K v'
and Neel surface anisotropy constant K s :
Nanogranular Layered Magnetic Films 201

Heft
HeN == 4nM
41TM - 2~v - 4~s +. (6.6)

Fitting the experimental plot of HeN Heft (t- 1 ) at tt--1I < t2'l
( t -1) t;1 to Eq. (6.1) (6. 1) and
considering M~1. 9 kG for C08oFe20'
CoaoFezo, we get the anisotropy parameters Kv~
- 4. 4 x 10 6 erg/cm 3
-4.4 3
and K ss ~ 1. 6 erg/ erg/cmcm 2z . The rather high surface
anisotropy indicates formation of sharply separated metal layers, which
correlates with the presence of 5 pronounced Bragg peaks in the low-angle
X-ray diffraction pattern for t = = 20 A sample (Fig. 6.3). 6. 3). Note that if Eq. (6. (6.6)
6)
were extrapolated below t 2z ,, H eN would change its sign at t~
Heft t ~ 12 A, which is
similar to the typical values in continuous magnetic/non-magnetic multilayered
films (Farle, 1998). The negative K vv value is most likely due to in-plane
constrictive strains produced in the metal layers by the rigid Ab Aiz 0 3
3 spacers.
These features, together with the observation of almost saturated magnetic
remanence for this thickness region, permits us to identify its magnetic state
as FM phase.
Another type of thickness dependence is observed at t < t l ' where
already Heft HeN (( t) is found linear. This can be compared with the linear
dependence of Heft HeN in 3D granular films on volumetric filling factor fv(Pogorelov
et ai.,
al. , 1999b), related to the volume density of the magnetic moment. In 20
granular layers the same role should be played by the surface density of
magnetic moments, which is proportional to the nominal thickness t.
However, the standard Langevin fit of the magnetization curve for the t = 10 = loAA
sample gives the effective SPM moment ....... 10 4 J.Js J.JB (Bohr magneton), while the

expected average moments of individual granules of size d ....... 3 nm (as seen


d.......
from HRTEM micrographs) are only""'" 1033 J.Js. J.JB' Hence the magnetic response at

. room temperature and in this thickness range is rather due to correlated


moments, in groups of""'" 10 or more granules. Such correlations between still
disconnected granules can be produced by dipolar stray fields at appropriate
granule shapes and distances between them. Thereafter, the main contribution
HeN is provided just by the correlated moments, and the observed linear
to Heft
law ....... (t - t' ) with t' ~6. 5 A is related to their partial surface density. The
SPM phase extends up to the magnetic percolation point t 11 , above which the
infinite cluster of correlated moments starts to grow at the expense of finite
SPM clusters, defining the transition to a mixed (M = SPM + SFM) phase.
Within this phase, structural clusters also (continuous metal islands) appear,
giving rise to Neel surface anisotropy and thus reducing Heft HeN (compared to the
SP linear law), until the structural percolation is reached at t z2 and a
continuous FM phase sets in.
The thickness dependencies of parallel FMR linewidth are shown in
Fig. 6. 12. At room temperature, t:.H fJ.H (t) presents a pronounced peak near
t 1 , similar to that recently observed in CoFe-Al
CoFe-Ab03 z0 3 cermet films (Golub et al. ,
2001). It can be explained by the static fluctuations of random internal fields
Hi on granules, reaching a maximum at the magnetic percolation point. The
simplest estimate of this contribution at t< t 1 is given by the dipolar field on a
202 G. N. Kakazei et al.

cluster produced by its nearest neighbor cluster, giVing ferromagnetic


t * )3 /2, which qualitatively fits the observed
IlH - ( t - t'
resonance linewidth t:.H
t:.H(t) at t<t 1 and contrasts with a linear behavior IlH(x)
IlH(t) t:.H(x) at low volumetric
filling factor x of magnetic metal component for 3D cermets. For t> t 1 , Hi
should mainly fluctuate due to relatively rare
rare""holes"
holes" in the FM background,
which are only healed well above t 2 2 where IlH
t:.H falls down to a small (almost
temperature independent) value for continuous film.

250
0\ - 0 - !'>.H,
!'>.H r at 77 K
o\
0\ -.-!'>.H,
-.-!'>.Hr at 300 K K
200

f='
E
o\o
? b

,,\
150
"5

~
.~

.... \
<lJ
C
:= 100
e<:
2:
u...

50 I." \
I .,,~~
.,,~0..
II
(I "'~--O
""~--O
I ----------.
-----------.
oL-":-g
'---8-!:---1:"::2----:'-16,-------::'20,--------,-J24~
--j:-'::2----::'-16:----='20:---:-'24-=-
CoFe CA.)
CaFe layer thickness (A)

Figure 6. 12 Linewidth of parallel FMR in [CaFe AI, 0 3 (30 A) ] 10 multilayers vs.


[CoFe ( t) / Ab
nominal thickness t of CaFe
CoFe layers at room temperature and at 77 K.

At 77 K, the IlH t) curve behaves quite differently below t 1: instead of


t:.H ((t)
decreasing, it continues to grow with decreasing t. This suggests that at low
temperature some additional mechanism of inhomogeneous broadening enters
the picture below t 1.l ' Such mechanism can be provided by the blocking of

cluster magnetic moments in non-collinear directions to the external field. This


blocking for FMR processes with time duration _10- 9 s, sufficient to realize rf
absorption by a magnetic moment in a metastable state, should be much more
effective than for static magnetization measurements needing -1 s. Also this
blocking can be enhanced by a high density of metastable states for a cluster
moment when all neighboring clusters are confined to the film plane, and when
the frustrated moments of single granules with appropriate aspect ratios are frozen.

6.6
6. 6 Transport and Magnetotransport Properties

The electrical resistance R was measured as a function of COaOFe20


COSOFe20 layer
(100 A~
thickness (1 t~25 A), temperature T(
A::( t::(25 T ( 10- 300 K) and magnetic field H
Nanogranular Layered Magnetic Films 203

(0 - 10 kOe) for the as-deposited samples in the CIP geometry (Kakazei


et al. , 1999). The CIP resistance measurements were made using in line four-
probe contacts, with silver paste or pressure contacts on the Al z0 3 top layer.
AI 2
The voltage probes were -- O. 65 cm apart, near the center of 2. 5 cm x
2.5 cm samples. We also measured the MR ratio, MR = R (H, T) / R (0, T) -
1, and the I-V characteristics both at room temperature and at low
temperature (20 K).

6.6.1 Electric Conduction in Zero Magnetic Field


The room temperature electrical resistance decreases sharply with the
Fe20 layer thickness, from 16 MQ at t = loA to 22 Q for t =
increase of Coso Fezo =
25 Ac
A(Fig.
Fig. 6. 13). The last resistance figure is consistent with the estimated
value for a stack of ten Coso Fezo
Fe20 continuous layers, each 25 A A thick. Such
behavior suggests a change from the dielectric to the metallic regime. This is
directly seen in the temperature dependence of the resistance, changing from
tunnel (dR/dT<O) at t~18t<18 AA (Fig. 6. 14a) to the metallic behavior (dR /dT>O)
above 18 A A of Coso Fezo
Fe20 thickness (Fig. 6. 14b), confirming the percolation
thickness previously referred. However, a slight upturn is still observed at low
enough temperatures 50 50 K, Fig. 6. 14b), indicating the presence of a
remanent tunnel contribution, likely associated with few tunnel bridges
between continuous parts of a metallic network within the CosoFezo
CosoFe2o layers.
T=300K
100 1=16 A

~ 50
1/'
1/'
l/r=20K
//T=20K
c: 0 ,j/
~.
01----~-1--1----
t
8(5 -50
/-1 .
/ //j
-100//
-100//
-4 -2 0 2 4
Voltage (V)

10 12 14 16 18 20 22
COSOFe20 layer thickness (A)

Figure 6. 13 CIP resistance of DMIM vs. VS. metal layer thickness. Inset: Typical I-V
characteristics obtained for the 16 A sample at room temperature and at 20 K.

A quantitative analysis has been done on the thermally activated R ( T)


dependence. As shown in the inset of Fig. 6. 14a for the t = 16 A
A sample, the
data agrees well with the standard dependence 10gR logR = 2 ( C / k BB T ) 1/2
I/Z
204 G. N. Kakazei et a!.
al.

/
4.0

~ 3.5
c:
3.0
o 0.1 0.2 0.3
I/TI/1/2
1/T
20
oO!:---:-!l100
0::c0----::200
2:-:!-0.". .0----::3:7
0"""0
300
Temperature (K)
(a)

1=20"\
t=20"\

o 100 200 300


Temperature (K)
(b)

Figure 6. 14 Temperature dependence of electrical resistance at different CaFe


CoFe layer
thicknesses. (a) At t < 18 A activated behavior is extended over the whole
temperature range. Inset: a linear InR vs r- 11/2
/
2
plot permits us to determine the
activation energy C """ O. 5 meV. (b) At t> 18 A activated behavior only holds at
C:::::;
lowest temperatures, changing for metallic behavior at higher temperatures.

characteristic of the low electric field limit for tunneling (Sheng et al. , 1973),
where C is the activation energy. This is expected to occur in a granular layer
Ab 0 3 ,, since the potential drop at each tunnel bridge in a series of--
CoFel Ab03 of""'" 10 6
bridges per cm is extremely small (see below). Correspondingly, the 1- V
characteristics should be linear for all the samples in the whole temperature
range as actually observed (inset to Fig. 6.14). 6. 14). Also, electrical breakdown is
virtually excluded in our films, unlike in tunnel junctions where the external
voltage is applied to a single bridge.
For our samples with t < 18 A, we obtain C ranging from:::::::::; from::=:::::: O. 5 meV at
t= 16 Acinset of Fig. 6. 14a) to--8 to ....... 8 meV for t= 12 A. In cermet systems, the
C value was used to estimate the intergrain distance-to-size ratio sl d, from
the Sheng-Abeles relation (Sheng et ai., 1973): C :::::::::;8~e2 ::=::::::8~e2 (sl d)2 I (1 ( 1 + 2s1 d) ,
with the inverse decay length for tunneling ~ ....... -- 1 A
A-I.
-1. However, in our DMIM

samples such an estimate gives unreasonably small values sl d -- ....... 10- 22


,, since
for typical values of d -- ....... 30 A(concluded from the micrograph, Fig. 6. 4) we
obtain s less than an atomic spacing, which has no physical meaning. This
discrepancy indicates that the above relation does not hold in DMIM systems.
We also prepared a sample with t = = 14 A
A for CPP measurements, again
exhibiting a dR I d T < 0 behavior consistent with thermally assisted tunnel tunneling.
ing.
Nanogranular Layered Magnetic Films 205

1/2
However the 10gR logR '"- T- T -1/2 law is not observed in CPP geometry. This
difference arises from the large voltage applied in CPP between consecutive
layers, leading to the high electrical field regime (Sheng et al., aI., 1973), with
nonlinear I-V characteristics. Under CPP conditions, Coulomb blockade was
observed in CoFe( 15 A)/Hf0 2 (40 A) and Co( 15 A)/Si02 (50 A) DMIMs at
temperatures below 40 - 50 K (Sankar et al. , 1997; Dieny et aI., al. , 1998), but
in our CoFe(
CoFe(14 14 A) / Ab03
A)/AI 2 0 3 (30 A) sample we did not find clear evidence on this
effect even at T = 10K
T= 10 K (a finite slope of d I / d V vs V still persists,
dl/dV where I is
persists,where
electrical current, V is electrical voltage). We then estimate the granule
charging energy E c '" O. 1 meV, from the activation energy C-
Ec-O. C'" 3 meV derived
from fits to the R ( n measurements for a t = 14 A sample, using the standard
( T)
theory (Sheng et aI., 1973). From such relatively low E c value one expects
Coulomb blockade effects in our system essentially restricted to very low
temperatures '" - 1 K.
The next step in transport studies consisted in magnetoresistance
experiments under various conditions of geometry, temperature and granular
layer thickness.

6.6.2
6. 6. 2 Room Temperature Magnetoresistance

The CIP magnetoresistance for all the samples measured at room temperature
and in magnetic fields up to 10 kOe is displayed in Fig. 6. 15. The onset of the
magnetoresistive response correlates well with the transition from metallic to
ing) conductance at t ~ 18 A. With further decreasing
(tunneling)
dielectric (tunnel
thickness t, MR significantly increases due to growing magnetic disorder,
reaching the greatest saturation value (at 10 kOe) in this series, MR max of

o 1=18...\
1=18,.\

-I

~ -2
0)
<I.)
u
c:
~
~ -3
'(h
.~

.,e
o -4

:2~
So
:2 -5

-6
--{j

-7
-7LL,-----_ _-':- -!- ~--___:_'::_
_ILl0 '--------'::-5-----'0,-----------=5-----:170
-10 -5 0 5 10
Magnetic field (kOe)

Figure 6. 15 Field dependencies of room temperature MR at different thicknesses


of ferromagnetic layer.
206 G. N. Kakazei et al.

= loA, which is, to our knowledge, the highest value observed in


6. 5at. % t =
DMIM at room temperature. However, since our series does not extend below
t = loA, this figure should not yet be considered the optimum thickness for
highest MR amplitude
ampl itude in this system. On the other hand, from our experience
on other series of samples with even lower t values and also from general
physical arguments, MR ultimately decreases due to the loss of granule spin
polarization at their shrinking size. It is thus important to note that MR reaches
its maximum well inside the superparamagnetic regime. The low field behavior
of the magnetoresistance shown in Fig. 6. 16 is characterized by the sensitivity
coefficient S = dMR/ dH I H-O'
H-.O. This quantity depends strongly on the Coso
COgO Fe20
layer thickness, and displays a fairly sharp maximum for an optimum thickness
t* ==13A
13 A (Fig. 6.16), reaching MR-MR-1% 1 % in 40 Oe. Notice that t* does not
correspond to maximum MR itself, though, as seen in Fig. 6. 17, there is a
t) curve at this point. Also, at t> t ** there appears a weak
break of the MR ((t)
(few Oe) hysteresis in the MR ( H) curves, but its quantitative relation to the
emerging magnetic order would require a more precise experimental basis.

25 Ii 1.0

20 o 0.8

V
0.6 g
o

:.'-\
"
0.2

,'. -"'-..
"'-..

o 10 12 14 16 18 20
COgOFe20 CA)
COSOFe20 layer thickness (1\)

Figure 6. 16 Critical behavior of the maximum sensitivity of MR to field and


normalized magnetization remanence vs. layer thickness. Solid squares are the
experimental values and the solid line is a guide for the eye. Open circles are
experimental from Figs. 6.7 and 6.9 and the dotted line is a ex: (t - t)
t * ) 12 fit.

In an attempt to further improve the MR properties of our samples, we


performed different annealing
anneal ing treatments under Ar, from 250'C
250C /10 min to
450C /60 min. But unlike the usual MR enhancement found in cermets and
450'C
DMIM, we observed a MR decrease, e. g. , MR lowered from 5. 2 % to 3 % in
glass/ AI 2 0 3 (30 A) /[ COgO
the glass/AI Coso Fe20 (13 A) / AI
Ab2 0 3 (30 A) ] 10 sample. This
behavior is consistent with the HRTEM data (Fig. 6. 4), showing that a
granular structure close to equilibrium is already present in our as-deposited
films, and contrasts with the results by other preparation techniques (Sankar
et ai., 1997 and 1998), where an almost continuous metallic network is
Nanogranular Layered Magnetic Films 207

35
6
30
v
~

" 25
0~
~

0~
~
25
:::: '"uc:
t
;)?.
'-;; 20
4
~
52E~ 15
'v;
=2
"0
.
15 _..0 - -_ _ e
~c:
;:z
;;Z bIJ
~ 10 2
:::,
~
"
~

o L..-,-':-_--,.J:-_---,L-_--'-:-_-"'lI1,--'
'---,-L--_-----'::-_ _"-:-_-'-:-_---'1I':c---' 0
12 14
Thickness of COSOFe20 layer (A)

Figure 6. 17 Comparison of thickness dependencies of room temperature MR ratio and


f::::,,13 A.
MR field sensitivity dMR/dH in CIP geometry, revealing anomalies close to 1",=,13

formed initially, requiring further annealing


anneal ing to obtain granules through
segregation.
Some explanations of this dissimilarity can be proposed. First, a stronger
non-wetting effect between CoFe and AI z 0 3 (larger than, e. g. , between Co
Al 2
SiOz2 or CoFe and Hf0
and Si0 HfO2z used in other works) favors the formation of the
granular structure in the very deposition process. Besides, the granule
segregation can be stimulated by a higher roughness of the underlying surface
of (amorphous) A1 20 3 , giving more nucleation centers for CoFe granules
A1z03'
(though
Cthough this hypothesis needs further experimental check). Starting already
with a granular structure, the main annealing effect might be oxidation of the
Coso Fe20 granules, with the subsequent decrease/loss of their spin
COBO Fezo
polarization.

6.6.3 Influence of the Number of Bilayers on Resistive and


Magnetoresistive Properties
The effect of the number of bilayers in multilayered film N on Rand MR was
studied in the Ab03C30
AIz03C30 A)/[CosoFe2o(13 A)/AI 20 3 C30 A)]N films. To assure
A)/[Co Bo Fe zo C13 A)/AIz03C30
the absence of electrical shortcutting between the layers by the current
contacts themselves, we repeated the measurements in CIP geometry with
three different types of contacts atop the sample: CD pressure golden
(Z) silver painted contacts, and @ deposited AI contacts. In all
contacts, (2)
cases, the experimental results were identical, with the zero magnetic field
resistance proportional to 1/ N, indicating the current to flow uniformly within
all the layers. This result agrees with the result previously obtained by
Morawe and Zabel and consistent with a sufficient interlayer conductance to
equilibrate potential across the planes.
208 G. N. Kakazei et al.

It was also found that, under an appl


applied
ied magnetic field, MR does not
depend on the number of bilayers, especially in the low/medium field range
(Fig. 6. 18). Some small changes of MR with N in high magnetic fields
( 5 %) are not systematic and are Iikely
(5%) likely due to slight
sl ight variations in the
deposition parameters for each film.

o - - N= I layer o
- - - N=2 layer
_I
-I N=8 layer

-4
-) ':-00:"70--5"'c:L
-I 0-=-0-'-0--""'
-=0-'-0-'-0-""50'5-0':-0""10-0-'
-07-'-10""0-'
0""0-0
Magnetic field (Oe)

-6
-10 -5 0 5 10
Magnetic field (kOe)

AI 20 33(30
Figure 6. 18 Dependence of MR in Ab A) /[Co so Fe20 (13 A) / Ab03
/[0080 Ah03 (30 A) IN on
the number of layers.

There are two important conclusions from this independence. The first,
physical conclusion is on weak magnetic correlations between neighbor
magnetic layers compared to the in-plane correlations. This supports our
suggestion arising from the analysis of the difference in CIP-MR and CPP-MR
temperature dependencies (see below). It can be understood comparing
magneto-dipolar interactions between much closer and better coordinated
granules in the same layer and between more distant and vertically misaligned
granules from adjacent layers. We also arrived at this conclusion from the
ferromagnetic resonance (FMR) studies on our films with different numbers of
ilayers, where the resonance field was found almost independent of N .
bilayers,
b
The second conclusion is technological, on the possibility to obtain the
same MR properties in a single magnetic layer Al AI z20 3 ao Fezo
3 /Co BO Al z0 3 structure
Fe20 / Ah03
as in a corresponding multilayer. Additionally, it is much easier to tailor the
properties of a single ferromagnetic layer. In fact, such a single layer presents
a 2D granular film. To optimize its MR performance, one should take into
account the following moments.
Directly deposited AI z0 3 provides the best known substrate (compared to
Al 2
SiO2
Si0 al. , 1998), Hf0
z , Dieny et aI., HfO2z (Sankar et aI., 1997), Zr02 ZrOz (Auric et al. ,
2000) to form a Co (CoFe) granular layer by room temperature deposition of a
proper amount of metal. However, a directly deposited AI z0 3 is not the best
Al 2
material to fill the intergrain spacers, if one seeks maximum MR values. It is
known from the studies on magnetic tunnel junctions that directly deposited
Nanogranular Layered Magnetic Films 209

Ab20 3 spacers give MR values of only ....., 2 % - 4 %, in contrast with ....., 40 %


AI
when Ab03 is formed by oxidation of preliminary deposited AI (Cardoso
et al., 1999). Though the latter route is difficult in granular media, it can be
viable in a 20 granular film. Besides, to fill the spacings and to cover the
metal granules formed on an AI 20 3 substrate, one might use a different
insulating material reliably deposited from a target.
glass/ Ab 0 3 ( t) /CO SO Fe20 (13 A) / AI
In our single layer granular films glass/Ab Ab20 3
(30 A) we observed sensitivity of their transport properties (R, MR) to the
Ab 0 3 buffer layer thickness t, both Rand MR decreasing as t increases
above 50 A. This can be due to the fact that the buffer layer surface becomes
smoother with growing b, resulting in fewer nucleation centers for CoFe
granules. Non-wetting may then lead to formation of larger metallmetallicic patches, a
suggestion that should be checked by structural characterization ( HRTEM,
AFM).

6.6.4 Temperature Dependence of Magnetoresistance


The MR(
MR ( H) curves under CIP
CI P geometry change only weakly with temperature
down to about 100 K. Below this temperature, which can be associated with
the blocking temperature T bb (as confirmed by SQUID measurements) one
observes the onset of hysteresis and coercivity for the 16, 14 and 13 A
samples, which are enhanced as temperature decreases. As illustrated in
= 15 K, hysteresis extends up to 2 kOe
Fig. 6. 19 for the t = 14 A sample at T =
and the coercive field He reaches 250 Oe. The hysteresis regime is

0 T=300K
T=300 K

-2

~-4

te
<l)
0
<.)
U
.;S0: T=15 K
T=15K
<f>
.~
.~ -2
-2
~
~
5J -4
51-4
~'"'"
::;E
Q..
d..
U -6

-8

-6 -4 -2 0 2 4 6
Magnetic field (kOe)

Figure 6. 19 Field dependencies for CIP-MR in Coao Fe20 (14 A) / Ab 0 3 (30 A) sample
at room temperature (upper panel) and low temperature (lower panel) .
210 G.N.
G. N. Kakazei et al.

accompanied by a significant increase in the saturation MR value. A similar


effect was recently observed in cermet Co-AI-O films and explained by higher-
order tunneling processes (Mitani et aI.,al. , 1998b).
Within our accessible experimental range, the highest observed MR
- 11 % occurs in the t = 13 A sample at 15 K, below which
saturation value of .....
the high resistance values make measurements increasingly more difficult.
Huge (GOhm-range) resistance values precluded low temperature MR
measurements in the tt< < 13 A samples.
The maximum room temperature MR value found in CPP geometry was
about 3 % in the sample with t = = 13 A. Note that MR considerably decreases
(by absolute value) with increasing bias voltage above ..... -2020 mV. A similar
tendency has been observed in spin junctions (Moodera et al., aI., 1995) and
Co/Si0 2 discontinuous multilayers
multi layers (Dieny et aI., 1998). The MR ratio at a
fixed current changes weakly with temperature in the whole range down to
10K. Also MR reveals hysteresis
hysteres is at low temperatures Iike
like in the CIP
CI P case,
within almost the same range and with similar coercive fields.
One interesting result is the much stronger temperature dependence of MR
for CIP than CPP geometry (Figs. 6. 19 and 6.20). To analyze this effect, we
consider the spin-dependent tunnel conductance G ..... - 1 + p2 cose (Inoue and
Maekawa, 1996), where P is the spin polarization of the ferromagnetic
material and eis the angle between magnetic moments of neighbor granules.
Assuming that tunnel processes in DMIMs are mostly in series for CIP and

0 T=300 K
-I
-[

-I

-2
~
~
0
~

c.::
~
~
cl..
c!... 0 T=15 K
0-
u

-I

-2

-8 -4 0 4 8
Magnetic field (kOe)

COgO Fezo ( 14 A) / Aiz 0 3 (30 A) at room


Figure 6. 20 Field dependencies for CPP-MR in Coso
temperature (upper panel) and low temperature (lower panel). Inset: bias voltage
dependence of CPP-MR.
Nanogranular Layered Magnetic Films 211

mostly in parallel for CPP, we obtain the respective (saturation) MR ratios:


= P2(1-
MR c,P = cose) ~ ) / (1 - p2<cose>~)
<cose>~)/(l-
p2 (1 - < p2 <cose) ~ ) (6.7)
= p2 (1 - <case>
MR cpp = case) z ) / (1 + p2)
p2).. (6. 8)
(6.8)
~.z are respectively for neighbor granules in the same layer
Here the averages < )>_L.z
and across layers, taken at H = 0 (or at H = He),He)' and the terms - p4 and higher
order are dropped in Eq. (6. 7). Of course, MR can increase with the polarization
p at lower temperatures but this effect should affect both geometries and it is
expected to be small. However an important temperature dependence can arise due
cose > depending on the particular geometry
to the magnetic correlation factor <cose)
(Dieny et al., 1998). At room temperature, < < cose )>~ is rather high due to the
strong in-plane dipolar coupling. But it should decrease considerably below the
blocking temperature T bb (when a strong short-range disorder is produced at
magnetization reversal near the coercive field, e. g., due to the increasing
anisotropy fields), leading to the increase of MRc,P up to- p2. From the measured
value- 9 % (Fig. 6. 17), we obtain P ~ O. 3, which is a reasonable figure for
COaoFe2o
COso Fe20 (Bardos, 1969). In contrast, the absence of such an increase in MRcpp is
likely due to the fact that a random correlation between granule moments in different
(almost uncoupled) magnetic layers is less sensitive to the change of ordering in
each layer with temperature.
It is known that an independent characterization of magnetic phase states,
in particular reflecting the short-range correlations between granule magnetic
moments, can be obtained from low field MR measurements. We thus
performed a detailed study of the temperature dependence MR ( n at a fixed
(T)
low field H = 45 Oe (it thus scales with the field sensitivity S ( T for a
selected sample with t = = 13 A, and the results are shown in Fig. 6. 21. The
observed maximum of MR (T) ( T) near ~ 250 K is, within experimental errors,
fairly close to the above determined SPM-SFM transition temperature T c for
this sample. For comparison, the inset of Fig. 6.21 6. 21 shows the thickness
dependence S ( t) measured at RT, with a sharp and symmetrical peak at tt'* ..
It is clearly seen that the maximum of MR ( T) is much broader and has a
sizeable asymmetry on the two sides from T c. The latter indicates the
difference in short-range magnetic correlations for SPM and SFM states. The
relation between the transport and magnetic properties of the considered
system can then be summarized as follows.
Direct metallic percolation among the CoaoFe2o grains needs layers as
thick as t"?t~ 18 A, leading to the onset of metallic conduction as indicated by
the change of R( n curve (Fig. 6. 14) and by the null TMR effect (Fig. 6.15).
R (T) 6. 15).
This is further supported by the qualitative change of the X-ray diffraction
pattern with appearance of extra Bragg peaks for t"? t ~ 18 A(
A (Fig.
Fig. 6. 3). But
ferromagnetic behavior starts to occur earlier, within yet discontinuous
magnetic layers for t> 13 A, Iikely
likely promoted by long-range dipolar
interactions. Near this critical thickness the magnetic susceptibility should
peak, thus producing a peak of MR sensitivity to field. On the other hand,
212 G. N. Kakazei et a!.
G.N. al.

E='
~
0.2~
0.8
'S
>
0.1 :~
c
"'"
<n
</>

0
12141618
Co Fe layer thickness (A)
CaFe

0.4

100 150 200 250 300


Temperature (K)

Figure 6. 21 Magnetoresistance of the 13 Asample in a fixed appl ied field H == 45 Oe as


a function of temperature (the points are experimental, the line is only a guide for the
eye),
eye). Estimated peak position (shown by the arrow) is about 250 K, close to the Curie
temperature T c """240 K for this sample determined in Figs.6.
Figs,6, 8 and 6.
6, 9,
g.

TMR is mainly due to granules in the SP fraction, which grows at decreasing


t * in the same critical way as the FM fraction goes to 0, giving rise to a
t - t'
certain break of MR(
MR (t)
t) (Fig.
(Fig, 6. 17). A further study of the effects of dipolar
interactions between disconnected magnetic granules (and related magnetic
correlations) is needed to conclude whether the discussed phenomena of
magnetic percolation correspond to a real thermodynamical phase transition,
transition.

6. 7 Conclusions

We discussed the various characteristics of nanogranular metal-insulator


layered systems which demonstrate their diverse properties, specific of this
composite structure and not existing in separate components.
components,
The most important physical feature is the observed possibility for
appearance of ferromagnetic order in a structurally disordered system of
dipolar coupled magnetic moments, which can be associated with a new kind
of phase transition occurring in the presence of long-ranged and frustrated
interactions and reduced dimensionality.
dimensionality, The difference of such ordering from
common Curie point behavior is seen in that it develops in an essentially non-
uniform way, similar to the onset of metallic behavior in various percolation
processes (Stauffer, 1979). However in this case the extending property is
the magnetic correlation between still separate metallic islands, which
Nanogranular Layered Magnetic Films 213

explains the usage of the term "magnetic percolation". This scenario is


interesting because of its similarity with such non-traditional types of magnetic
ordering as the Griffiths phases in randomly diluted magnets (Griffiths, 1969)
or phase separation in colossal magnetoresistance manganites (Uehara et al. ,
1999). As seen from the data in Section 6. 5, such transition in DMIM systems
can take place through variation of nominal thickness t at fixed temperature t
and with varying t at fixed t, defining a certain continuous line in the t- T
plane. The main practical importance of this fact is that magnetoresistance
exhibits a sharp peak of its field sensitivity at crossing the transition line, and
the physically non-accidental character of such singularity is confirmed by the
correlated singularities observed in other dynamical magnetic properties, as
the FMR linewidth. The peculiar non-uniform character of the emergent long-
range order is also reflected in the magnetic behavior at low temperatures
(below Curie point),
point) , where another, spin-glass-like transition can occur. This
looks similar to the known re-entrant transitions in mixed continuous magnets
(Srivastava et al. , 1994), but has the novel feature of co-existence between
ordered and glassy subsystems.
Another important practical conclusion is that nanogranular layered
unlike
systems, unl ike usual continuous magnetic layers, display almost the same
efficiency in single-layer as in multilayered structures. This can greatly simplify
preparation and will improve stability of multilayered devices, using simply a
single layer of nanogranules on a well characterized substrate with optimized
insulating covering. We hope that the forthcoming studies along these lines will
bring a significant progress in our understanding of essential physics and in
practical performance of nanogranular magnetic layered systems.

References
Auric, P., J. S. Micha, O. Proux, L. Giacomoni and J. R. Regnard. J.
Magn. Magn. Mater. 217: 175 (2000)
Babonneau, D. , F. Petroff, J. -L. Maurice, A. Vaures and A. Naudon. Appl.
Phys. Lett. 76: 2892 (2000)
Lett.76:
Baibich, M. N. , J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P.
Etienne, G. Creuzet, A. Friederich and J. Chazelas. Phys. Rev. Lett. 61:
2472 (1988)
Bardos, D. I. J. Appl. Phys. 40: 1371 (1969)
Bouchaud, J.P.andP.G. Zerah. Phys. Rev.B47: 9095 (1993)
Briatico, J. , J. -L. Maurice, J. Carrey, D. Imhoff, F. Petroff and A. Vaures.
Eur. Phys. J. D 9: 517 (1999)
Brown W.F., Jr. Phys. Rev. 130: 1677 (1963)
Cardoso de Freitas, V. Susana Isabel Pinheiro. PhD Thesis. Instituto Superior
Tecnico, Lisbon, Portugal (2001)
Desmicht, R. , G. Faini, V. Cros, S. F. Lee, A. Fert, F. Petroff and A.
Vaures. Appl. Phys. Lett. 72: 386 (1998)
214 G.N. Kakazei et al.

Dieny, B. , S. Sankar, M. R. McCartney, D. J. Smith, P. Bayle-Guillemaud


and A.E.
A. E. Berkowitz. J. Magn. Magn. Mater. 185: 283 (1998)
Evoy, S., D. W. Carr, L. Secaric, Y. Suzuki, J. M. Parpia and H. G.
Craighead. J. Appl. Phys. 87: 404 (2000)
Farle, M. Rep. Progr. Phys. 61: 755 (1998)
Fettar, F. J.-L. Maurice, F. Petroff, L.F. Schelp, A. Vaures and A. Fert.
Thin Solid Films 319: 120 (1998)
Gehanno, V., P. P. Freitas, A. Veloso, J. Ferreira, B. Almeida, J. B.
Sousa, A. Kling, J. C. Soares and M. F. da Silva. IEEE Trans. Magn. 35:
4361 (1999)
Giacomoni, L. Ph.DPh. 0 Thesis, Universite Joseph Fourier Grenoble. France
(1998)
( 1998)
Gittleman, J I., Y. Goldstein and S Bozowski. Phys. Rev. B 9: 3891
( 1972)
Golub,V.a.,
Golub,V.O., G.N. Kakazei,A.F. Kravets, N.A. Lesnik,Yu.G. Pogorelov,
J.B. SousaandA.Ya.
Sousa and A.Ya. Vovk. Mat. Sci. For.373-376: 197 (2001)
Griffiths, R. B. Phys. Rev. Lett. 23: 17 (1969)
Inoue, J. and S. Maekawa. Phys. Rev.BRev. B 53,11927
53, 11927 (1996)
Jay, J. Ph. , M. Wojcik and P. Panissod. Z. Phys. BIOI: 471 (1996)
Jonsson, T., P. Svedlindh and M. F. Hansen. Phys. Rev. Lett. 81:397681: 3976
(1998)
( 1998)
Kakazei, G. N. , A. M. L. Lopes, Yu. G. Pogorelov, J. A. M. Santos, J. B.
Sousa, P. P. Freitas, S. Cardoso and E. Snoeck. J. Appl. Phys. 87: 6328
(2000)
Kakazei, G. N.,
N. , P. P. Freitas, S. Cardoso, A. M. L. Lopes, M. M. Pereira
de Azevedo, Yu. G. Pogorelov and J. B. Sousa. IEEE Trans. Magn. 35:
2895 (1999)
Kakazei, G. N., Yu. G. Pogorelov,
Pogoreiov, A. M. L. Lopes, J. B. Sousa, P. P.
Freitas, S. Cardoso, M. M. Pereira do Azevedo and E. Snoeck. J. Appl.
Phys.90: 4944 (2001a)
Kakazei, G.N., Yu.G.
Yu. G. Pogorelov,
Pogoreiov, J.B.
J. B. Sousa, V.a.Golub,
V. O. Golub, N.A.Lesnik,
N. A. Lesnik,
S. Cardoso and P. P. Freitas. J. Magn. Magn. Mater. 226 - 230: 1865
(2001b)
Kaufman, H. R., R. S. Robinson and W. E. Hughes. Characteristics,
Capabilities and Applications of Broad-Beam Sources. Commonwealth
1987)
Scientific Corporation( 1987>
Kleemann, W. , Ch. Binek, a. O. Petracic, G. N. Kakazei, Yu. G. Pogorelov,
M. M. Pereira de Azevedo, J. B. Sousa and P. P. Freitas. J. Magn. Magn.
Mater. 226-230:
226 - 230: 1862 (2001a)
Kleemann, W. , a. O. Petracic, Ch. Binek, G. N. Kakazei, Yu. G. Pogorelov,
Pogoreiov,
J. B. Sousa, S. Cardoso and P. P. Freitas. Phys. Rev. B 63: 134423
(2001b)
Kumar, D. , J. Narayan, A. V. Kvit, A. K. Sharma and J. Sankar. J. Magn.
Magn. Mater. 232: 161 (2001)
Nanogranular Layered Magnetic Films 215

Luttinger, J. M. and L. Tisza. Phys. Rev. 70: 954 (1946)


Luttinger, J.M. and L. Tisza. Phys. Rev. 72: 257 (1947>
Maurice, J.-L., J. Briatico, J. Carrey, F. Petroff, L. F. Schelp and A.
Vaures. Phil. Mag. A 79: 2921 (1999)
Mitani, S., K. Takanashi, K. Yakushiji and H. Fujimori. J. Appl. Phys. 83:
6524 (1998a)
Mitani, S., S. Takahashi, K. Takanashi, K. Yakushiji, S. Maekawa and H.
Fujimori. Phys. Rev. Lett. 81: 2799 (1998b)
Miyazaki, T. and N. Tezuka. J. Magn. Magn. Mater. 139: L231 (1995).
Moodera, J. S. , L. R. Kinder, T. M. Wong and R. Meservey. Phys. Rev.
Lett. 74: 3273 (1995)
Morawe, Ch. and H. Zabel. J. Appl. Phys. 74: 1969 (1995)
Mosca, D. H., N. Mattoso, W. H. Schreiner, A. J. A. de Oliveira, W. A.
Ortiz, W.H.
W. H. Flores and S.R.
S. R. Teixeira. J. Magn. Magn. Mater. 231: 337
(2001)
Naudon, A., D. Babonneau, F. Petroff and A. Vaures. Thin Solid Films 319:
81 (1998)
Ogielski, A. T. Phys. Rev. 8 32.7384
Rev.B32. 7384 (1985)
Petracic, O. , W. Kleemann, Ch. Binek, G. N. Kakazei, Yu. G. Pogorelov,
J. B. Sousa, S. Cardoso and P. P. Freitas. Phase Transitions 75: 73
(2002)
Pogorelov, Yu. G., G. N. Kakazei, J. B. Sousa, A. F. Kravets, N. A.
Lesnik, M. M. Pereira de Azevedo, M. Malinowska and P. Panissod,Phys.
Rev. B860:
60: 12200 (1999a)
Pogorelov, Yu. G., G. N. Kakazei, M. M. Pereira de Azevedo and J. B.
Sousa. J. Magn. Magn. Mater. 196-197: 112 (1999b)
Prakash, S. and C.L. Henley. Phys. Rev. B 842:
42: 6574 (1990)
Sankar, S.,
S. , A. E. Berkowitz and D. J. Smith. Appl. Phys. Lett. 73: 535
( 1998)
Sankar, S., B. Dieny and A. E. Berkowitz. J. Appl. Phys. 81: 5512 (1997>
DienyandA.E.
Schaadt, D. M. , E. T. Yu, S. Sankar and A. E. Berkowitz. Appl. Phys. Lett.
74: 472 (1999)
Schelp, L. F. , E. L. Rosa, J. -L. Maurice and F. Petroff. J. Magn. Magn.
Mater. 205: 170 (1999)
Schelp, L. F. , A. Fert, F. Fettar, P. Holody, S. F. Lee, J. L. Maurice, F.
B 56, R5747 (1997>
Petroff and A. Vaures. Phys. Rev. 856,
Sheng, P. , B. Abeles and Y. Arie. Phys. Rev. Lett. 31: 44(44 ( 1973)
Sousa, J. B., G. N. Kakazei, Yu. G. Pogorelov, J. A. M. Santos, O.
Petracic, W. Kleemann, Ch. Binek, S. Cardoso, P. P. Freitas, M. M.
Pereira de Azevedo, N. A. Lesnik, M. Rokhlin and P. E. Wigen. IEEE
Trans. Magn. 37: 2200 (2001 a)
(2001a)
Sousa, J. B. , G. N. Kakazei, Yu. G. Pogorelov, P. P. Freitas, S. Cardoso,
A. M. L. Lopes, M. M. Pereira do Azevedo, J. A. M. Santos and E.
Snoeck: Mater. Sci. For.373-376:
For. 373-376: 81 (2001b)
216 G. N. Kakazei et al.

Sousa, J. B. , G. N. Kakazei, Yu. G. Pogorelov, S. Cardoso, P. P. Freitas,


A. M. L. Lopes and M. M. Pereira de Azevedo. In: P. Entel and D. E. Wolf
eds. The Structure and Dynamics of Heterogeneous Systems. World
Scientific, Singapore, p. 199(2000)
Srivastava, B. K., A. Krishnamurthy, V. Ghose, J. Mattson and P.
Nordblad. J. Magn. Magn. Mater. 132: 124 (1994)
Stauffer, D. Phys. Rep. 54: 1 (1979)
Takzei, G.A., I. Mirebeau, L.P.
L. P. Gun'ko, 1.1. Sych, O.B.
O. B. Surzhenko, S.
V. Cherepov and Yu. N. Troschenkov. J. Magn. Magn. Mater. 202: 376
(1999)
Thomson, T. , C. Riedi, S. Sankar and A. E. Berkowitz. J. Appl. Phys. 81.
5549 (1997)
Uehara, M., S. Mori, C. H. Chen and S. -W. Cheong. Nature 399: 560
( 1999)
Wojcik, M., J. P. Jay, P. Panissod, E. J~dryka, J. Dekoster and G.
Langouche. Z. Phys.B 103: 5 (1997)
Zanghi, D., A. Traverse, F. Petroff, J-L. Maurice, A. Vaures and J. P.
Dallas. J. Appl. Phys. 89: 6329 (2001)

It is our pleasant duty to acknowledge the important contributions to many of the presented
colleagues, P. E. Wigen (Ohio
results by our colleagues. COhio State University,
University. Columbus), E. Snoeck
(CEMES-CNRS, Kleemann, Ch. Binek and O. Petracic (Gerhard-Mercator
CCEMES-CNRS, Toulouse), W. Kleemann.
University, Duisburg), N. A. Lesnik and V. O. Golub (Institute of Magnetism.
Magnetism, Kiev), P.
Panissod (IPCMS,
CIPCMS, Strasbourg). M. Pereira de Azevedo and A. M. L. Lopes (University
<University of
Funda<;ao de Cielncia
Porto). Also the support from Portuguese Fundacao Ciencia e Tecnologia under the
projects 3/3. 1/MMA/1787 /95./95, 2/2. 1/FIS/302/94, and P/CTM/11242/98 is gratefully
acknowledged.
7 Monodisperse Ferromagnetic Metal Particles:
Synthesis by Chemical Routes, Size Control and
Magnetic Characterizations

Philippe Toneguzzo, Guillaume Viau and Fernand Fievet

7. 1 Introduction

The physico-chemical properties of well-defined metal particles ranging from a


few nanometers to several micrometers make them of interest for numerous
appl ications in different fields such as catalysis, electronics, optics or
applications
medicine. This is particularly true in the areas of magnetism and magnetic
materials since magnetic properties are strongly dependent on particle shape,
size, dispersion, elemental composition and surface oxidation. Thus, during
the last few years, nanoscale ferromagnetic metal particles have been the
object of increasing interest:
( 1) to investigate size and surface effects upon magnetic anisotropy
(Respaud et al. , 1998) and coercitivity (Gangopadhyay et aI.,
al. , 1994) ;
(2) to understand the mechanism of magnetization reversal in single
particles (Wernsdorfer et al. , 1997> ;
(3) to study collective properties of long-range ordered assemblies of
monodisperse particles (Legrand et al. , 2000).
Fine metal ferromagnetic particles in the nanometer and submicrometer
size ranges also have different applications as advanced materials, including
particulate recording devices for information storage, ferrofluids, and
microwave composite materials. For this last application, ferromagnetic
metal-based materials display properties that make them of interest as higher
and broader working frequencies than bulk ferrimagnetic oxides. As far as
microwave absorbing properties are concerned, metals have to be used as
fine particles dispersed in an insulated matrix. Such composite materials
exhibit magnetic losses in the microwave frequency range due to the
gyromagnetic resonance within the particles. Dimensions of the metallic
particles are kept below the skin depth, meanly -10 IJm in the GHz range. In
the last decade numerous works studied how the microwave properties of a
composite depend on the intrinsic properties of the magnetic particles, on their
volume concentration and their ordering. Two main objectives can be defined:
first, the design of high permeability composite materials with an optimal
218 Philippe Toneguzzo et a!.
al.

control of permeability level, resonance frequency and resonance width;


second, a better understanding of the dynamic magnetic properties of fine
particles and a tentative correlation with their morphological characteristics
(size and shape) and their static magnetic properties. From an experimental
point of view, in both cases, the control of the morphology of the
ferromagnetic particles is needed. Magnetic resonance is highly dependent on
the particle shape through the effect of the demagnetizing field. So, materials
made up of particles with poorly defined shapes present a very broad
resonance band. According to theoretical studies, the magnetic resonance was
also expected to depend on the particle size in the submicrometer and
nanometer size ranges. An accurate determination of size effects required
model materials made up with calibrated particles.
Hence, for use as advanced materials in high technology or as model
materials in fundamental studies, non-agglomerated, uniform metal
ferromagnetic particles with a controlled mean size, a narrow size distribution
and a definite alloy composition are required. Different methods including
physical, chemical and electrochemical ones have been used in order to yield
uniform metal powders. Among them, chemical precipitation from
homogeneous solutions, if it can be conducted under controlled conditions,
appears as a very convenient route. In the first section, the guidelines, which
allow us to control the particles size in such precipitation processes in order to
obtain monodisperse system, will be recalled. A particular emphasis will be
laid on the polyol process as a versatile method, which allows us to make
quasi-spherical non-agglomerated particles for both Co-Ni and Fe-Co-Ni
systems with various compositions. An accurate control over particle size in
the nanometer to micrometer range and over the particles size distribution can
be achieved by this process. In the second section, the morphological,
structural, textural, chemical and magnetic characteristics of such powders
will be described and compared with those of powders obtained by other
chemical or physical routes. It will be shown in the last section how polyol
made particles have led to a better understanding of the dynamic magnetic
properties of fine particles through the comparison between experimental data
obtained on well-defined materials and theoretical calculations.

7. 2 Preparation of Monodisperse Fine Ferromagnetic


Metal Particles by Chemical Routes

Synthesis of ferromagnetic metal particles by precipitation from solution is of


particular interest since it allows us:
( 1) to control the particle characteristics (shape, average size, particles
size distribution) by acting upon experimental parameters such as
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 219

concentration of reactants, temperature, addition of nucleative or protective


agents;
(2) to provide finely dispersed metals in rather large amount with particle
ranging from nanometer to micrometer size;
(3) to yield single metal or alloyed particles as well.
Neutral metal atoms can be provided from dissolved species either
through reduction of Co (II), Ni(II),
Co(II), Ni (II), Fe (II/III) ions or complexes or through
Fe(II/III)
decomposition of zero-valent precursors such as metal carbonyl or
organometall ic complexes.
organometallic

7. 2. 1 Mechanism of Particle Formation


Whatever the reaction involved, the precipitation of the metallic phase occurs
in two steps: nucleation and particle growth. Since the metal atoms are highly
insoluble, as soon as they are generated they form, by a stepwise addition,
thermodynamically unstable clusters called embryos. Embryos are formed and
dissociate into atoms according to a dynamical process. As further metal
atoms are generated, embryos become larger and lead to stable nuclei when
they reach a critical size. As nucleation results from local statistical
fluctuations, this step is favored by a high supersaturation of metal atoms and
then by a high rate of formation of metal by reduction or decomposition of the
dissolved species. From these nuclei the particle's, growth may proceed at
first by a stepwise addition of new metal atoms and lead to nanosized primary
particles. Particle growth can continue either by the further addition of metal
atoms controlled by their diffusion toward the solid surface or by the
coalescence of the primary particles and formation of larger secondary ones.
According to this mechanism of particle formation, as far as metal
nanoparticles are desired, precipitation must be limited to the early stages of
the growth step leading to non-agglomerated particles in the nanometer size
range. Stable colloidal dispersions are obtained either through steric
stabilization with protective agents such as polymers or surfactants or through
electrostatic stabilization. Larger particles in the submicrometer or micrometer
size ranges can be obtained by further growth. If the particle growth occurs by
the stepwise addition of metal atoms, a general condition must be fulfilled in
order to obtain non-agglomerated metal particles with a well defined shape,
a controlled average size and a narrow size distribution: nucleation and growth
must be two completely separated steps (Overbeek, 1982). Spontaneous
nucleation occurs when the concentration of the metal generated by the
reduction reaches a critical supersaturation level. If the nucleation rate is high
enough, then the sudden nucleation lowers almost immediately the
concentration below this critical supersaturation level. If the rate of formation
of the metal is not too high this nucleation step is followed by the growth of the
particles from the original nuclei without formation of new nuclei as long as the
220 Philippe Toneguzzo et al.

metal is slowly generated provided that coalescence is prevented. On the


contrary, if the particle growth occurs by coalescence of primary particles,
secondary particles with irregular shape and large size distribution would be
expected. In fact, it has been observed.
observed, at first for systems such as oxides or
hydroxides,
hydroxides. that such a growth mechanism can yield secondary polycrystalline
particles with a uniform shape and a narrow size distribution (Ocana et al al..,
1995; Jezequel et al. , 1995). This can be also observed for metals (Goia and
Matijevic.
Matijevic, 1998) including ferromagnetic ones (Viau et al. , 1996b; Toneguzzo
al ., 2000) with the formation of mostly spherical particles.
et al.

7. 2. 2 Control of the Morphological Characteristics of the Particles

7.2.2.
7.2. 2. 1 Nanoparticles

The final size of the particles depends on the nucleation step and the growth
step as well. For a given amount of metal obtained from decomposition or
reduction of the precursors this size depends on the number of nuclei generated
during the nucleation step. Higher is this number, finer are the particles.
Consequently, to prepare nanosized dispersion, a high supersaturation level is
required to generate a large number of nuclei. A high supersaturation ratio can
be reached if the rate of formation of metal atoms is high. Such a condition can
be fulfilled by different ways regarding the reaction involved.
If the precipitation of metal is carried out through reduction, a high rate is
favored by the use of a strong reducing agent acting upon uncomplexed species
and by a high reaction temperature. Thus, monodisperse nanocrystals in the
size range 2 - 11 nm are obtained from reduction of CoCb by lithium
borohydride (LiBEt 3 H) in dioctylether at 200'C 200C (Sun and Murray,
Murray. 1999).
Similarly, ultra fine iron powders are synthesized by reduction of FeCb
Similarly. FeCl z with an
organometall
organometallicic complex of V ( II) namely V (CsHs)z (CsH s ) 2 in tetrahydrofuran
(Choukroun et al. , 1998). The crystallites
crystall ites average size was evaluated to be
18 nm. The preparation of stable colloids dispersion or nanostructured
assemblies require also the control of the growth step by the addition of
stabilizers. Co nanoparticles obtained by Sun and Murray are made in the
trialkyl phosphines (PR 3 ) Short chain alkyl
presence of oleic acid and trialky!
phosphines (R = = n-C4 Hg ) allowed fast growth favoring production of larger
particles (7 - 11 nm) than those (2 - 6 nm) obtained with a reduced growth
=
rate observed with longer chains (R = n-Cs H17 ). ) . Similarly, Choukroun et al.

controlled the growth of Fe particles by the addition of polyvinylpyrrolidone


(PVP) they obtained smaller particles (ca. 8 nm instead of 18 nm without
PVP) agglomerated into 50 - 200 nm superstructures with a surprising common
magnetic and crystallographic orientation.
The control of the growth of nanosized ferromagnetic metal particles can
be achieved efficiently when the reduction is carried out in microemulsions
Monodisperse Ferromagnetic Metal Particles: Synthesis by...
by. . . 221

rather than in homogeneous solutions. Thus, inverse micelles, i. e. , water in


oil droplets stabilized by a monolayer of surfactants, have been used as
microreactors to obtain Co nanoparticles using sodium borohydride NaBH 4 as
a reducing agent (Petit et al. , 1997, 1998; Chen et al. , 1994). The diameter
of the droplets is controlled by the volume of water dissolved. Petit et al.
mixed two micellar solutions, one containing cobalt bis (2-ethylhexyl)
(2-ethylhexyl )
sulfosuccinate Co (AOT)2
( AOT ) 2 and the other Na (AOT)
( AOT) and NaBH 4. Exchange
between droplets occurs as they collide
coli ide due to Brownian motion, hence the
reduction yields Co particles with an average diameter estimated to 6. 4 nm
with a polydispersity of 21 %. CoCI 2 2 is used as precursor and the particles

diameter is varied from 1. 8 to 4. 4 nm by varying the CoCh concentration


(Chen et al.,
aI., 1994). For the same system, the particle size is controlled in
the range 3. 8 - 8. 8 nm by the so-called germ-growth method where particles
obtained in a first stage are used as nucleation sites to grow larger particles
(Lin et aI.,
al., 1999).
Ferromagnetic metal nanoparticles can also be obtained by decomposition
of zero-valence precursors such as metal carbonyl Fe (CO) 5' CO 2 2 (CO) 8 ,
Ni(CO)4 or metal nitrosyl carbonyl Co(NO)
Co (NO) (CO)3. These precursors are
dissolved in organic solvents, and the fast nucleation followed by a rapid
quenching of the particle growth can be achieved by ultrasonic irradiation of
the solution. The chemical effect of the ultrasound originates from the
formation, growth and subsequent implosive collapse of bubbles in the liquid
(Grinstaff et aI.,
al., 1993). This phenomenon called acoustic cavitation
generates transient spots with high local temperature and pressure followed by
a very rapid cooling which favors the formation of amorphous metal
nanoparticles. Iron (Grinstaff et al.,
aI., 1993; Cao et al., aI., 1995), nickel
(Koltypin et aI., 1996), cobalt (Shafi et aI., 1998a), Co-Ni (Shafi et al. ,
1998b), and Fe-Ni-Co (Shafi et al. , 2000) alloys amorphous powders have
been obtained by this sonochemical method. Stabilizing agents have to be
added to obtain non-agglomerated particles. It must be pointed out that it is
difficult to obtain particles with a narrow size distribution and a tailored
average size by such a method. When the sonication is carried out with a low
power apparatus, it is noteworthy that crystalline powders are obtained
instead of amorphous ones. Hence Fe particles are obtained (mean radius 1,
(90.1 nm); standard deviation of size distribution a= (0. 550. 05) nm)
with a bcc structure for the smallest particles and a fcc structure for the largest
ones (de Caro et al., 1996). Suitable experimental conditions have been
found recently to obtain a fast nucleation step followed by a controlled particle
growth during the rapid pyrolysis of CO2 2 (CO) 8 (Puntes et al. , 2001). These

authors produced monodisperse, defect-free, spherical nanocrystals whose


size can be controlled and varied between 3 and 17 nm by controlling
parameters such as precursor/surfactant ratio, reaction temperature and
injection time.
Another approach to achieve a fast nucleation providing metal
222 Philippe Toneguzzo et al.

nanoparticies
nanoparticles is the decomposition of organometallic precursors, namely
alkene complexes. The decomposition of such a precursor dissolved in an
organic solvent is carried out under dihydrogen in the presence of a stabilizing
polymer at room or moderate temperature. Non-agglomerated Co and Ni
nanoparticles with a mean diameter in the range 1 - 2 nm and a very narrow
size distribution have been obtained by this method (Osuna et al., 1996;
Respaud et al. , 1998; Ould Ely et al. , 1999).

7. 2. 2. 2 Micrometer and Submicrometer-sized Particles: the Polyol


PolyoI Process
Ferromagnetic metal particles can be obtained in this size range from reduction
of metallic species in homogeneous solutions. The most straightforward route
to obtain well-defined particles with a uniform shape, a narrow size distribution
and a low agglomeration is to find, as far as possible, suitable experimental
conditions where the nucleation appears as an initial burst of nuclei followed by
a diffusional growth of the particles without significant coalescence of these
primary particles during this growth step. The number of nuclei generated
during the nucleation step must be low with respect to the synthesis of
nanoparticles. Hence, the rate of formation of metal atoms must be low
enough to achieve such a control of the nucleation. Moreover, this slow rate
allows an easy control of the growth rate as well, without formation of new
nuclei during this step. Such a kinetic control of nucleation and growth of metal
particles requires the use of mild reducing agents.
1. General mechanism
Liquid polyols, namely ex-diols such as 1, 2-ethane diol (ethylene glycol:
Em
EG) or 1, 2-propanediol (propylene glycol: PrEG) are used in the so-called
polyol process (Viau et al., aI., 1994, 1996a, 1996b; Toneguzzo et al. , 2000;
Fievet et aI.,al. , 1989a, 1989b, 2000). Such liquid polyols are hydrogen-bonded
liquids that are able to dissolve, to some extent,
extent. ionic inorganic compounds
due to their rather high relative permitivity. Moreover, they are mild reducing
agents that are able to reduce to zero-valence state ions or complexes of noble
metals or copper at room or moderate temperature but also Co ( II ]I )) or Ni ( II
]I )
species at temperature up to their boiling point (close to 190'C 190C under
atmospheric pressure). Polyols are not able to reduce directly Fe ( II ]I ) or Fe
( ill) ions to the zero-valence state, nevertheless iron or iron-based particles
can be obtained through the disproportionation of Fe ( II) ]I) hydroxide in such
mild reducing media. They allow us to avoid the competitive oxidation of Fe
( II]I ) into Fe( ill) that is observed in water (Shipko et al. , 1956). Moreover,
they act as a complexing agent to retain Fe ( ill) resulting from the
disproportionation as dissolved complexes and to prevent the crystallization
crystall ization of
magnetite. Therefore, the simultaneous precipitation of magnetite and iron is
prevented. Polyols molecules act also in some extent as protective agents by
adsorption on the metal particles. Hence, they allow us to prevent or at least
to control the coalescence of ferromagnetic metal particles during their growth
and then to obtain well-defined and non-agglomerated particles. Polyols are of
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 223

particular interest as reacting media due to these various functions. Indeed,


they act as solvent, reducing agent, crystal growth medium and complexing
agent for the metallic cations. In contrast to other chemical methods, no
further reducing agent has to be added to the solvent to achieve the formation
of metal particles. Co, Ni, Fe powders made up of well-defined particles, but
also polymetallic powders such as Co-Ni, Fe-Ni and Fe-Co-Ni have been
obtained by the polyol process under experimental conditions which have been
detailed previously (Viau et al., a!., 1996a, 1996b; Toneguzzo et al.,
ai., 2000;
Fievet 2000). In all cases the synthesis of mono or polymetallic powders is
achieved according to the following scheme:
( 1) dissolution of the metal salts with a suitable amount of sodium
hydroxide in the polyol;
(2) addition of a solution of a nucleating agent if heterogeneous nucleation
is required;
(3) heating of the solution and precipitation in situ of an intermediate
phase (metal alkoxide hydroxide);
(4) progressive dissolution at higher temperature of this slightly soluble
precipitate which liberates
Iiberates in solution solvated cationic species;
(5) reduction by the polyol or disproportionation of these metal cations;
(6) nucleation and growth of the metal ferromagnetic particles from the
solution.
The role of the intermediate phase must be emphasized. When the
reaction occurs, the concentration of the cationic species in solution is
controlled by the progressive dissolution of this sparingly soluble phase acting
as a kind of reservoir. The dissolution equilibrium regulates the release of
these species, controls the supersaturation ratio and then allows having a very
short nucleation step. Nevertheless, it is possible to have an important yield of
metal at the end of the growth despite the low supersaturation ratio.
2. Heterogeneous nucleation
In some cases, to obtain more easily the separation between the
nucleation and growth steps homogeneous (spontaneous), nucleation can be
replaced by heterogeneous nucleation by adding a suitable nucleating agent.
Salt or complexes of noble metals (NM) such as Ag, Pd or Pt are used. Those
metallic species are readily reduced to the zero-valence state by the polyol
and yield metal nuclei prior the precipitation of the ferromagnetic metal.
Whereas the mean size of Co, Ni and Co-Ni particles obtained by
spontaneous nucleation liesIies in the micrometer range, submicrometer-sized and
nanometer-sized Co-Ni and Fe-Co-Ni ones are synthesized by heterogeneous
nucleation. Moreover, by varying the nature and the concentration of the
nucleating agent, it is possible to control the number of nuclei and therefore to
have an accurate and reproducible control of the particle size in this range
(Fig. 7. 1). Indeed, the particle's mean size steadily decreases whereas the
[NM J/ ([ Fe J + [Co]
molar ratio [NM]/([Fe] [Co J + [Ni])
[Ni J) increases whatever chemical
composition and noble metal used. This variation is explained according to the
224 Philippe Toneguzzo et al.

following assumptions:
( 1) the size of formed nuclei does not depend on the amount of the noble
metal initially introduced;
(2) each nucleus yields one single ferromagnetic particle through a
diffusional growth without significant coalescence of primary particles.

(a)

100 11111
I1Ill

(b)

Figure 7. 1 TEM image of C050 Ni 50 particles made by the polyol process through
heterogeneous nucleation using Pt as nucleating agent: (1) d m = 110 nm([Pt]/[Co+
2. 5 x
Nj) ] = 2
ND] X 10- =
10 - 55 ) ; (b) d mm = 50 nm
nm(( [[Pt]/[ =
Nj) ] = 2 5 xX 10-
Pt] / [ Co + ND] =
10 - 44 ) ; (c) d mm = 6 nm
X 10- 2 ).
([Pt]/[Co+ Nj)] = 2.55x
([Pt]/[Co+ND]=2
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 225

This model yields a linear relationship between the mean diameter d m and
[NMJ/([FeJ + [Co]
the inverse of the cubic root of the molar ratio [NM]/([Fe] [CoJ + [Ni])
[NiJ)
according to the following equation:

d = (6M
(6MFecoNi V FecoNi
VNMPNM) X NM PNM ) X (( [NM]
[NMJ ))-1/3
-1/3 (7. 1)
0.1)
m npFeCoNi M NM
NM M [Fe]
[FeJ + [Co]
[CoJ + [Ni]
[NiJ
where MFecoNi. M NM , PFeCoNi' PNM' are the molecular weight and the density of
MFeCoNi' M
the polymetallic
polymetall ic FeCoNi powder and noble metal, metal. respectively.
respectively, and V NM NM the
average volume of one nucleus of the noble metal. If the mean diameter d m
([NMJ/([FeJ + [Co]
appraised from image analysis is plotted against ([NM]/([Fe] [CoJ +
[NiJ) ) -1/3 for different compositions in both CoNi and FeCoNi systems with
[Ni]
different nucleating agents (Ag. (Ag, Pt), the expected Iinearlinear relationships.
relationships,
according to Eq. 7. 1. 1, are obtained except for the seeding with Ag at high
concentration and for the COzo C020 Ni so composition for the seeding with Pt (Fig.
7.2). This different behavior between Pt and Ag and this discrepancy with the
model observed for the highest silver concentrations are tentatively ascribed to
the stronger tendency of the silver nuclei to coalesce. The size of the noble
metal nuclei therefore becomes concentration dependent (Viau et al. , 2001).
In fact.
fact, Pt and Ag appear as complementary nucleative agents. For instance
nm~dm~
CosoNbo' Pt is rather used to synthesize the finest particles (25 nm:S;;;dm:S;;;
for CosoNizo.
150 nm) whereas Ag is preferred for the larger ones (200 nm:S;;;d nm~dm~500
m:S;;;500 nm)
400
t:. a
6.
b
o c
x d
300

t:.
6.

I 200
1200
t:.
6.

oJ
"<:sE

o 50 100 150
l3
([NM]/[Fe]+[Co]+[Ni]r I/3/
([NM]I[Fe]+[Co]+[Ni]r

Figure 7. 2 Variation of the particle's mean diameter d m against the inverse of the
cubic root of the molar ratio [NMJ/([FeJ + [CoJ + [NiJ) for different compositions in
both CoNi and FeCoNi systems and for Pt and Ag as nucleating agents: (a) C0
Coso
80 Nbo Ag:
Ag;
Coso Ni so Pt; (c) C080 Nbo Pt; (d)
(b) C050Ni50Pt:(C)CosoNboPt; Feo 13 [COso Ni so Jo 87 Pt.
(d)Feo13[C050Ni5oJos7PI.
226 Philippe Toneguzzo et al.

(Toneguzzo et al. , 1998). Furthermore, the average size of the noble metal
nucleus is obtained through the slope of the linear relationships exemplified on
Fig. 7.2, the mean size of the Pt nuclei.
nuclei, i. e. , the kernel of the ferromagnetic
metal particles finally obtained is typically found in the range 2 - 3 nm.
Thus, monodisperse Cox Ni 1oo
Thus. - x powders can be obtained by the polyol
100 -

process in the whole composition range including pure Co or Ni powders. The


particle size can be varied over a wide range,
range. which can reach three orders of
Fez[CoxNilOo-x]l-z powders have been obtained as
magnitude. Monodisperse Fez[CoxNi100-x]1-z
well, but in this case the iron content is Iimited
limited by disproportionation, hence
the composition of the powders could be varied in the ranges O~z~O. 25 and
O~ x ~ 100. In fact,
O~X fact. the mechanism of growth of these particles prepared in
polyols depends on the nature of the metal considered (Viau et al., al . 1996a).
The morphological and textural characteristics of Cox Ni 100 ,oo - x particles (cf.
section 7.3.4.
7. 3.4. 1) are consistent with the diffusional growth mechanism while
Ni 1 - z ones are consistent with a growth mechanism involving
those of Fez Ni,-z
coalescence of primary particles. For Fez[CoxNiloo-x]l-z
Fez[CoxNi,oo-x]'-z it is likely that the
mechanism involves both diffusional growth and coalescence.
Whatever the growth mechanism.
mechanism, the reduction or disproportionation of
metallic compounds in liquid polyols appears as a versatile method of
preparation of ferromagnetic metal powders made up of non-agglomerated
particles with a well-defined shape and a narrow size distribution. It allows an
accurate control of the mean size of the particles in the submicrometer and
nanometer size ranges and of the composition of polymetallic powders. The
morphological, textural and structural characteristics of the obtained particles
will be described in the next section.

7. 3 Characterization of Ferromagnetic Metal Particles


Made by the Polyol Process

7.3. 1 Morphological Characteristics


The control of the morphological characteristics of ferromagnetic metal
particles is of particular interest since magnetic properties are strongly
dependent on particle shape, size and dispersion. For most appl ications and
theoretical studies as well, non-agglomerated particles with a definite shape.
shape,
an accurately controlled mean size and a narrow size distribution are
preferred.

7.3. 1. 1 Particles Shape


For industrial uses, in particular for powder metallurgy requirements, the
by. . .
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 227

spherical shape is generally preferred. Thus, industrial powders such as


micrometer-sized iron powders made either by decomposition of Fe (CO) 5 in
the gaseous state (Mond Langer process, Antonsen and Tundermann, 1984) or
by atomization under nitrogen (Fig. 7.3), 7. 3), show a very low deviation from
sphericityCD. Forfiner iron particles, i. e. , inthenanometer-sizerange,
ironparticles, in the nanometer -size range, obtained

(a)

20,------------------,
20

15

o 1.5 3.5 5.5 7.5


~.~ ~,P1 ~
9.5 11.5
1\.5 13.5
d(~1Jn)
d(J.1ITI)
(b)

Figure 7.3 (a) SEM image of iron-based particles made by atomization under neutral
gas( d m = 4. 5 lJm, a = 2. 6 lJm);
IJm, (j IJm); (b) histogram of size distribution and (c) histogram of
deviation from sphericity.

CD The deviation from sphericity S is defined from MEB or TEM image analysis as S =
4lTAP-2z with A the area of the 20
4lTAP- 2D image of the particle and P its perimeter.
228 Philippe Toneguzzo et al.

1.00

0.75 f ------
-- -------
-- -- --
-- ---- - { 0

~ 0.50
0.501--~---------__+
E
::J
o
U

0
0.25. 2 5 r - - - - - - - - - - - - - - j /

o 0.955 0.965
0965 0.975 0.985 0.995
Sphericity
(c)

Figure 7.3 Continued from page 227.

by the vapor deposition method (Gangopadhyay et al., aI., 1993), the deviation
from sphericity raises as the particles mean diameter increases with a
maximum geometric ratio always lower than 2. Otherwise, for specific
applications such as magnetic recording, an acicular shape for the magnetic
fillers is preferred because of the higher coercive field induced by the shape
anisotropy. For this purpose, superfine Fe-Co and Fe whiskers with a long
axis in the range O. 1 - 1 IJm
lJm and remarkably low size and shape dispersions
are industrially manufactured (Hisano and Saito, 1998).
Equiaxed ferromagnetic metal particles are obtained by the polyol
process excepted for Fe-Co particles that are aggregates without definite

(a)

Figure 7. 4 (a) SEM image of polyol made Feo Feo. 13 [Co ao Ni 20 Jo


[C080 a7 particles ( d m =
Jo. 87
205 nm, a = 25 nm) ; (b) histogram of size distribution and (c) histogram of deviation
from sphericity.
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 229

40

30

10

o f/1
f'/l
O. J 4
0.14 0.18
.~m
0.22 0.26 0.30 0.34
dd(~m)
(11m)
(b)
100

75

~ 50

U
=
c:::l
'"
o

25

o I,...,
".., n
".., ,,,..,
,,..., ~,~
n,,..., "..,,~
,...",..., ".."".., n
,...",..., 1'1

0.79 0.82 0.85 0.88 0.91 0.94 0.97 1.00


Sphericity
(c)

Figure 7. 4 Continued from page 228.

shape. The deviation from a spherical shape is very small indeed for FeCoNi
particles (S == 0.99) whereas it is slightly larger for the CoNi ones (S =
= 0.80)
(Figs. 7. 4c and 7. 5c).

7.3. 1. 2 Particles Size


When their mean size increases from a few nanometers to several tens
nanometers, ferromagnetic particles behave at room temperature as
superparamagnetic particles, then as single-domain particles. In this size
range, a rather good control of the particle size can be achieved by different
230 Philippe Toneguzzo et al.

physical methods such as hydrogen plasma metal reduction (HPMR) (Li


et al. , 1997, 1998, 2000) or vapor deposition under an inert atmosphere by
acting upon suitable experimental parameters. By this last method for
instance, Fe and Co particles on one hand.
hand, Ni particles on the other hand.
hand,
have been made with a size lying in the range 5 - 35 nm and 9 - 85 nm
respectively by varying the inert gas pressure (Gangopadhyay et al.. aI., 1993;
Du et al. , 1991). Electrochemical
Electrochem ical or chemical
chem ical routes allow the control of the
size in the nanometer range as well (Sun and Murray,Murray. 1999; Kneller and
Luborski. 1963). Nevertheless,
Luborski, Nevertheless. the polyol process is of particular interest
since it allows an accurate and reproducible control in an extended size range
lying from a few nanometers (through heterogeneous nucleation with the most

I 250
25011111
11m ~
,

k_-_--
1..-_-_--
(a)

2
255,,--
----
----
--- -
----
-------------
----
---,

20 fJ-----------t:/)--r/..j----------1
- - - - - - - - - - - f / . . l - - r/ . 1 - - -

15 1 - - - - - - - 1 /..I--t'/
151-------(/. ~'-r/)___y/ . 1 - - - - -

e~...
~

]
1$
E 101-------1/
10 f----------Y/.J-f/J--j/
/JI----V/J--Y/}--f'7}
/}--f'7}---
::l
:::l
;z
;Z

5 J--V/.I----y;..-r-V/l-----t:;/J--V/H'/'J--t'>'1--

0.037 0.041 0.045 0.049


d(l-lIn)
d (1l111)
(b)

Figure 7.5 TEM image of CO'5 = 42 nm, a == 5 nm) ; (b) histogram of


Ni,5 particles (d m =
COGS Ni,s
size distribtion and (c) histogram of deviation from sphericity.
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 231

16
\6

12

~
E8
::l
o
U

o ~
0.5
~~ ~
0.6 0.7 0.8
~~~
0.9 1.00
Sphericity
(c)

7. 5
Figure 7.5 Continued from page 230.

concentrated Pt solution) to a few micrometers (through spontaneous


Co xNi lOo - xand Fez[CoxNilOo-x]l-z
nucleation) for CoxNilOo-xand Fez[CoxNiloo-x]l-z (cf. section 7.2.2.2).
7. 2. 2. 2).
Optimization of the magnetic properties involves that the standard
deviation of size distribution (j a must be as narrow as possible. The standard
deviation is generally lower for the chemical routes than for the physical ones.
As an example (j a is markedly higher than 100 % d m for particles obtained by
vapor deposition under an inert atmosphere (Gangopadhyay et al., 1993)
while (ja is close to 7 % d m for nanoparticules obtained by reduction in solution
using lithium borohydride as reducing agent (Sun and Murray, 1999).
The histogram of size distribution of polyol made particles (Fig. 7. 4b and
Fig. 7. 5b) roughly displays a gaussian shape, i. e., a symmetric layout,
rather than a log-normal one (generally observed with physical methods or
other chemical ones). The standard deviation is widely lower than 15% 15 % d m ,
and even lower than 10% 10%d d m for Fez[CoxNilOo-x]l-z samples.

7.3. 1. 3 Dispersion
Polyol-made ferromagnetic metal powders are made up of non-agglomerated
particles. Adsorbed polyol acts as a protective agent to prevent the
coalescence of the particles and to ensure a good dispersion. The formation of
chain-like arrangements due to dipolar interactions between neighboring
particles is only observed by TEM with the finest powders.
232 Philippe Toneguzzo et al.

7.3.2 Crystal Structure

Generally, ferromagnetic metal fine particles are crystalline whatever their


size and whatever the physical or chemical synthesis route except for the
sonochemical method which yields amorphous metal due to the rapid cooling
rate (> 109 K/ s) associated to acoustic cavitation (Grinstaff et al., ai., 1993;
Cao et aI.,al. , 1995; Koltypin et al. , 1996; Shafi et al., al. , 1998a, 1998b, 2000).
Metal powders made by the polyol process appear as crystalline as well,
metal phases being the only ones detected by X-ray diffraction (Fig. 7.6). For
single metal powders the expected crystal structure is evidenced, i. e. , body
centered cubic (bcc) and face centered cubic (fcc) for Fe and Ni
respectively. Cobalt has long been known to have two crystal structures:
hexagonal (hcp) close packed stable at room temperature and fcc stable
above 450 C. 'C. Nevertheless it must be pointed out that a complex cubic
structure (e-Co) isomorphous to ~-Mn has been evidenced in nanoparticles
obtained by chemical routes (Sun and Murray, 1999; Puntes et al., aI., 2001;
o inega and Bawendi, 1998). The two close packed polymorphs are present in
Dinega
polyol-made cobalt samples. For the polyol-made polymetallic powders, the
X-ray patterns depend on composition. In the Fe-Co system three phases are
shown identical to those expected for a mixture of the two metals (bcc for o:Fe exFe
and hcp and fcc for the two Co polymorphs). This observation is consistent
with the segregation of the metals inferred from X-ray microanalysis (cf.
section 7. 3. 3. 2). At the opposite end, in the Fez Ni 100 100 -- zz system a single fcc
phase is evidenced in the available composition range (z<20). (z <20). A fcc phase is
always observed either as a single phase or as the main phase in both
CoxNi100-x
Co x Ni 100 - x and FeOI3[CoxNiI00-xJ087
Fea13[Co x Ni lOO - x ]087 systems, a second hcp phase with weak and
broad lines appears for a cobalt content xx~35 ~ 35 and, for Feo FeO,1313 [Cox Ni 100
lOO -- xx Jo 87
]087

powders, a third bcc phase is indexed when x>80 (Toneguzzo et al. ,2000).
The accurate determination of the lattice parameter of the fcc phase shows a
linear dependence against composition over the whole metal composition
range in the Cox Ni 100 100 - lOO -- zz (z <
- xx and Fez Ni 100 < 20) systems. Such a linear
dependence, which therefore agrees with the Vegard law, has been reported
earlier for bulk CoNi (Taylor, 1950) and FeNi (Hansen et al., aI., 1985) alloys
and proves the existence of a fcc solid solution over the whole composition
range for the polyol made powders. In the Cox Ni 100 lOO -- xx system, the range of
existence of the hcp phase for Co-rich compositions is larger than that
predicted in bulk materials by the phase equilibrium diagram (Landolt-
Bornstein, 1982) since the hcp phase is only expected for x ~ 70 instead of
xx~35.
~35. In a similar way, FeNi nanoparticles elaborated by HPMR are
constituted by a mixture of bcc and fcc phases in a range of composition wider
than that in the bulk Fe-Ni alloys (Dong et ai., aI., 1999).
It must be stressed that the crystal structure appears size-dependent for
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 233

III
111 fcc
fcc
220
22 fcc
fcc
100
100hep
hcp 200 iCe
101 hcp fcc
II0
1I0hcp
ilep

Ill iCc

200 iCc
fcc

40 60 80 100
28 C
CO )(Co Ka radiation)
(a)

III fcc
fcc
002 hcp
hcp

220fcc
fec
110
II0hep
hcp

40 60 80 100
28 C
CO )(Co Ka radiation)
(b)

Figure 7.6 XRD patterns (Co Ka radiation)of


Kcx radiation) of powders of different compositions with an
almost constant particles size: (a) Cox Ni 100
lOo -- xx ;; (b) Feo 13 [COx Ni 100
lOo -- xJo
x Jo 87
87.
234 Philippe Toneguzzo et al.

polyol-made Co-Ni nanoparticles. The intensities of the fcc lines progressively


decrease with decreasing size and they almost vanish for the finest powders
= 5 nm) (Viau et al.
(d m =5 al.,, 2001). Such nanocrystals with a hcp structure are
tentatively considered as representative of the early stages of the growth of
the larger particles. Since the precipitation is faster for Co than for Ni, the
formation of Co-rich crystallites occurs during these early stages and their high
Co content favors a hcp close packing of metal atoms. Size effect on the
crystal phase has been previously reported for Fe (de Care
Caro et al. , 1996) and
Co nanoparticles. However for these latter particles obtained by physical
ways, the stabilization of the fcc structure has been evidenced for the finest
particles (Kitami et al. , 1997).

7.3.3
7. 3. 3 Chemical Composition

The optimization of static or dynamic magnetic properties of ferromagnetic


metal particles involves the control of the non-metal impurity level and for
polymetallic powders the control of the overall metal composition on one hand,
and of inter- and intra-particular composition homogeneity on the other hand.
These different points will be considered for the polyol-made powders with
respect to those obtained by other synthesis routes.

7.3.3.
7. 3. 3. 1 Overall Metal Composition
Polymetallic powders are obtained by chemical and physical routes that
involve at least one step with a homogeneous phase (gas or solution) in order
to achieve the mixture of the metallic precursors at a molecular level. The
overall metal composition of the as-made powders may be different with
regards to the nominal precursors composition. Thus, for instance in the
sonochemical preparation of CoNi amorphous alloys from Co Co(NO)
(NO) (CO)
(CO)33 and
Ni (Co) 4' (Shafi et al. , 1998b) the yield of decomposition is less for the
former and it is necessary to start with a well controlled excess of
Co(NOHC0)3
Co( NO) (Co) 3 to obtain the required alloy composition. In physical
processes, the chemical composition of the polymetallic ultrafine particles is
also generally different with regards to the nominal precursors composition.
Thus, the iron content of Fe-Ni nanoparticles obtained by HPMR is always
higher as compared to the nominal composition by about 2at. % - 3at. 3a1. % (Li
et al., 1998) while the composition of Fe-Co nanoparticles appears quite
identical to the nominal one. For the polymetallic powders made by chemical
routes from solution, the overall metal composition generally depends on the
ability of the respective cationic species to be reduced by the reducing agent
and on the existence of competitive reactions occurring during synthesis.
Hence, the reduction of Co and Ni species by polyols proceeds through very
similar mechanisms with a difference between the reduction rates small enough
to allow to vary the composition of the Cox
CoxNi lOO - x over the whole range. This
Ni lOo
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 235

composition can be controlled accurately since it is close to the nominal


composition of the solution of precursors in the polyol. At the opposite end,
Fez [Cox Ni 100 - x J1- z polyol-made powders show a Fe/M molar ratio (M=Co,
Fez[CoxNilOo-x]l-zpolyol-made (M = Co,
Ni) significantly lower than the ratio of the corresponding cations in the starting
confi rms that the yield of formation of iron is limited
solution. This result confirms lim ited by
disproportionation of Fe( II ) whereas Co( II) II ) and Ni ( II)
II ) are on the contrary
quantitatively reduced (Viau et al. a!. , 1996b). Therefore z can be only varied in
the range 0 - O. 25 whereas x can be varied in the range 0 - 100.
Non-ferromagnetic metals may be detected as impurities in ferromagnetic
metal fine powders elaborated by physical ways if they are present in the
precursors. In chemical
chem ical processes, the occurrence of such non-ferromagnetic
metal impurities depends on the process itself, the power of the reducing
agent and on the chemical purity of the precursors. Thus, X-ray fluorescence
analysis (XRF) on both FeCoNi and CoNi polyol made powders shows that the
contents of non-ferromagnetic metal impurities such as Na, Ca, AI are very
low (a few ppm), whatever the particle's size and composition. These
different elements are in the reacting medium in high content for Na (sodium
hydroxide concentration ranging from O. 25 to 1 mol/L) or low content
0. 1 %) for the others (impurities present in the starting precursors). The
cations of these electropositive metals are not reduced by the polyol owing to
its mild reducing power, and therefore they remain in the reacting medium.

7.3.3.2 Inter and Intra-particular Homogeneity in Composition


For a given sample of a polyol-made polymetallic powder the metal ratio
measured by X-ray microanalysis on several particles appears quite identical
to the overall composition measured by XRF or by Inductive Coupled Plasma-
Atomic Emission Spectroscopy (ICP-AES). Then, a good interparticular
homogeneity in composition is ascertained for both Cox Ni 100 1oo -- x and
Fez [Co x Ni lOo - x Jl- z powders. Moreover, X-ray microanalysis executed by
Fez[CoxNilOO-x]'-z
fixing the nanoprobe at the core and at the edge on one isolated particle shows
different elementary compositions for both Cox Fez [Cox Ni lOo - x Jl- z
Ni lOo - x and Fez[CoxNilOO-x]l-z
CoxNi
powders. Whatever particle's size and chemical composition in CoNi and
FeCoNi systems, the core of the particles is always richer in Co than the
edge, whereas the edge is richer in Ni. Conversely, Fe appears
homogeneously distributed within each particle (Toneguzzo et al. , 2000). The
presence of this composition gradient in each particle is explained by a slightly
higher rate of reduction for Co than for Ni. This difference is nevertheless
sufficiently low to prevent a segregation of the two metals in each particle. On
the contrary, for the Fez Co 100 - z particles, which are non-spherical and
agglomerated (cf. section 7. 3. 1. 1), 1 ), elementary distribution maps clearly
show that the particles are made up of tiny iron particles embedded in a cobalt
matrix (Viau et al. a!. , 1996a).
236 Philippe Toneguzzo et al.

7.3.3.3
7.3. 3. 3 Non-Metallic Impurities
Fine ferromagnetic metal powders obtained by physical ways are very air
sensitive. Usually, the particles are passivated for instance by the residual
oxygen present in the chamber during evaporation or by soaking during a few
hours in a dilute air-argon mixture prior to being slowly recovered under the
room atmosphere. Therefore, comb combined
ined oxygen and to a lower extent,
adsorbed oxygen are detected as the main non-metallic impurity. The latter is
present as absorbed water mainly due to air moisture. Combined oxygen
appears in an oxide or oxyhydroxide layer as inferred from various methods
such as XRD, electron diffraction or X-ray photoelectron spectroscopy (XPS). (XPS) .
Thus, Fe 3 3 0 4 and y-Fe 2 0 3 are the main phases evidenced with Fe particles but
Y-Fe203
FeO and FeOOH have also been reported (Gangopadhyay et al., 1991,
1992a; Uchikoshi et al., 1989) while CoFe 2 2 0 4 ferrite have been evidenced
with Fe-Co ultrafine particles (Gangopadhyay et al., al. , 1994).
Oxygen is also present as metal oxide in ultrafine powders synthesized by
chemical ways. Otherwise, the use of sodium borohydride or hypophosphite
as reducing agents in solution generates alloyed M x 8 B 1 - x or M x P1 - x particle's
with possible Band8 and P contents as high as 20at% (Oppengard et al., 1961;
Wells et al.,
aI., 1989). On the contrary, reducing agents such as hydrazine
(Sapiezko and Matijevic, 1980) and hydroorganoboranes (Bonneman (80nneman et al. ,
1990) do not generate such non-metallic
non-metall ic impurities. Carbon and hydrogen are
also generally detected in powders obtained by chemical routes since these
powders are obtained in organic solvents. These organic impurities are also
located on the particle's surface, but their nature seems to be very different
according to the process, to the reacting medium (solvents, precursors,
additives, etc.) and to the experimental procedures (drying, washing,
recovering, storage, etc.). The main non-metallic impurities in the polyol po/yoI
made particles are also oxygen, carbon and hydrogen (Toneguzzo et al. ,
2000). Their presence is expected owing to the experimental conditions,
i. e. , synthesis in an organic boil
boiling
ing solvent (1,
(1 ,2-propanediol),
2-propanediol), washing with
water, ethanol and acetone and drying under argon prior to recovering and
storage under air. The oxygen content is typically 1wt. 1wI. % - 1. 5wt. % and
4wt. % for as-made powders in Co-Ni and Fe-Co-Ni systems respectively. The
carbon and hydrogen contents can be lowered significantly by thermal
treatment under a neutral atmosphere at 350'C. The carbon content decreases
typically from 1. 5wt. % for as-made iron based particles to O. 15wt. % -
O. 35wt. % after annealing, and even from 1wt. % to lower than O. 15wt. % for
Co-Ni compositions. For both systems, the carbon and oxygen contents
increase steadily when the mean particle size decreases. Then it may be
supposed that these impurities are mainly located at the surface of the
particles. This point will be discussed in detail in the next section.
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 237

7. 3. 4
7.3.4 Core-Shell Microstructure
In finely div~ded ferromagnetic metal powders, physical properties, such as
saturation magnetization and density, differ from bulk values and vary with the thl
particle'
particle'ss size. A continuous decrease of saturation magnetization of
ferromagnetic particles against their size has been often reported in literature
(Haneda, 1987; Yamada et a!., al., 1985; Gangopadhyay, 1992b; Li et aI., a!. ,
1994; Hadjipanayis et a!.al. , 1991). For instance, the saturation magnetization
of Fe particles prepared by the evaporation-deposition technique decreases
from 190 to 25 emu/gemu/ g for a mean diameter ranging from 20 to 6 nm
(Gangopadhyay et al., aI., 1992a, 1992b, 1993; Hadjipanayis et al., 1991; Li
et al., 1994). The decrease of saturation magnetization with decreasing
particles size is related to a core-shell microstructure, i. e., a metallic
metall ic core
with an oxide outer layer. The higher surface to volume ratio in the finest
particles results in a much higher contribution of this oxide layer (constituted of
Fe3 0 4 and/or y-Fe2 0 3 ).
). The magnetization values smaller than 80 emu/ emu/g g
(bulk saturation magnetization for Fe oxides) observed experimentally for the
finest particles may be due to the existence of nonmagnetic surface layers
(dead layers, Berkowitz et al., aI., 1968) around the particles or due to spin
canting in the outer shell (Gangopadhyay et al., 1993; Haneda and Morrish,
1989).
1989) .
For CoNi and FeCoNi polyol-made particles, the density and saturation
magnetization decrease steadily with decreasing size while C, 0 and H
impurities contents increase steadily (Toneguzzo et al., 2000). All these
behaviors are also consistent with the core-shell model. Accordingly, it has
been assumed that all the particles of a given sample are spherical with the
same radius (reasonable assumptions owing to the measured sphericity and
the narrowness of size distribution) and are made up of a core which is mainly
(but a priori not exclusively) constituted by ferromagnetic metals and of a thin
coating layer of impurities. Then, the increase of the overall impurities
contents, the decrease of the density and of the saturation magnetization with
decreasing particles size are explained by the increase of the relative volume
fraction of the less dense and less magnetic superficial layer assuming that its
thickness is constant. From these hypotheses, the following equations can be
obtained after limited development by assuming that the layer thickness e is
significantly lower than the particles radius R and that the saturation
magnetization of the layer M Ms,'
s . , is equal to 0:

Tp -=
- Tc + 33ee PI
~(Tl
- ( Tl -- Tc R- 1
))R- (7.2a)
Pc
Pp = Pc-
Pc - 3e(pc - PI)R- 1
3e (Pc-PI)R- (7.2b)
238 Philippe Toneguzzo et al.

(7.2c)

where T,P,
T, p, V and M ss are elemental impurity content (wt%)
(wt %) for instance C
content, density, volume and saturation magnetization respectively, the
subscripts p, c and I being related to the whole particle, the core and the
impurity layer respectively.
If the experimental T p' PP and M ss.p. p are plotted versus the reciprocal of

the mean radius of the particles R as inferred from the Eqs. (7. 2a) - (7. 2c)
(Fig. 7.7
7. 7 to Fig. 7. 9), the expected Iinear
linear relationships are obtained with a
good coefficient of linear regression. It is shown in Eqs. (7. 2a) - (7. 2c) that
the impurities contents, the density and the saturation magnetization of the
core can be inferred from the intercept of the different lines with the ordinate.
The different values obtained from this model for different compositions in both
CoxNi,oo-xand Feo 13[CoxNi100-xJo87
CoxNilOo-xand ,3[CoxNi,00-xJ087 systems are given in Table 7.1.
99 ~--~---~----~---~
~--~---~---~---~ 3000

Oxygen
...
Carbon
0o Hydrogen

6 2000
~

~
~ E0-
Q.
(5 .3:
2:
c.5 :t:",0-
::r::
,..f- <.P-
3 1000

o 0.02 0.04 0.06


R I (limn)
1
(l/n111)

Figure 7.7 (wt. %) vs. the reciprocal of the particle's mean


7. 7 C, 0 and H contents <wt.
composition( Co
radius R for a given composition< BO Ni zo
Coso 20 )
) ..

7.3. 4. 1 Core Characterization


For Cox Ni lOO - x powders, the core density and saturation magnetization are
NilQo-
close to the bulk values while for Feo.,3 FeO.13 [Cox Ni 100
,oo - x J087 system, they are
significantly lower than the bulk ones. In both cases they decrease steadily as
the cobalt content increases. These characteristics for the CoNi particles are
consistent with an almost non-porous core formed by ferromagnetic metals with
a very low content of non-metall
non-metallic ic phases and a low porosity. On the contrary
in the Feo 13 [Cox Ni ,oo
loo - x Jo 87 system, the lower core density and saturation
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 239

magnetization with respect to the bulk cannot be explained only by the core
impurities (C and 0)
Q) contents that are higher than in the Co-Ni system. These
characteristics must also be related to a closed porosity within the metal core.
Such conclusions are supported by the size of the crystallites estimated from
dark-field transmission electron microscopy (TEM) observation on one hand
and from X-ray diffraction (XRD) lines broadening analysis on the other hand
9
X Co zo Ni 80
CozoNi so
o CosoNi
Coso Ni so
Co 8o Ni zo
.A. CosoNi
.II.
6. Feo I3[Co
I:; 80 Ni 20 ]o 87
I3[CosoNizo]o S7
FeO.I3[CosoNiso]087
Feo.dCosoNiso]OS7
8 1---.----'l''I2'>~---+----+-------j
1-------'l'~=------t-------+-------1

6 "--_ _-----'-_ _----->-------'--- --'---'''''---_->-..J


o 0.02 0.04 0.06 0.08
R- 1 (limn)

Figure 7.8 Real density p of C020 Ni 80 C050 Ni 50


ao ,Coso C080
so ,Co Nbo , Feo
ao Ni,o 13 [C0
[Coso
50 Ni 50 Jo. 87
so Jo a7 and
Feo. 13 [C0
Feo ao Nbo
[Co 80 Ni 20 Jo. a7 powders vs. the reciprocal of particles mean radius R.
Jo 87

160
160 r - - - - - - , - - - - - - - - - - - - ,
Co 8o Ni zo
CosoNi
o CosoNi so
x Feo dCosoNiso]o
I3[CosoNiso]OS7
87
135 .A. Co
.II. zo Ni 80
CozoNi so

'OIl
Oil
::;
"":3
E 110 k=--n;:;--+---+---~-+----f
~
0-

~'"
'.:;;.':i

851------+-----+----""'......,-+------1
85

60
60 L -_ _---l.-_ _---lL-_=-'--_ _..,---,J
o0 0.02 0.04 0.06 0.08
R- 1 (lImn)
(I/nm)

Figure 7. 9 Saturation magnetization Ms of C020 Ni80 C050 Ni 50


ao ' Coso so ', C0
COao Nbo and
80 Ni,o

Feo 13 [C0 50 Ni 50
[COso so Jo a7 powders vs. the reciprocal of particles mean radius R.
Jo. 87
240 Philippe Toneguzzo et al.

(Toneguzzo et al. , 2000). Mean crystallite sizes of iron-based powders are


always significantly smaller than for the CoNi ones and the crystallites sizes
obtained through the Williamson and Hall analysis (Williamson and Hall, 1953)
are consistent in both cases with the sizes obtained from dark-field TEM
images. Iron-based and particularly FeNi particles (Viau et al., 1996b) are
constituted by many equiaxed crystallites whose mean size is a few
nanometers, while CoxNi lOo - x particles are made up of some larger crystallites
CoxNilDD-x
with a more irregular shape (Toneguzzo et al. , 2000). It has been suggested
(Viau et al. , 1996b) that these different characteristics can be related to two
(cf. section 7.2.2.2).
distinct growth mechanisms (ct. 7. 2. 2. 2).

Table 7. 1 Density p, M,. impurities (carbon and oxygen) contents


p. saturation magnetization M"
T of the particles core of different compositions in the Cox
CoxNilOO-x
Ni IOO - ' and
Feo 13 [Cox Ni loo - x Jo a7 systems. The experimental values are inferred from Eqs.
FeoI3[CoxNiI00-,J087
(7. 2a) - (7. 2c) and compared with bulk data.

pc Pbulk M s c
.c Ms,bulk
MS,bUlk Ms.c
M,.c T c,carbon
c.carbon T c,oxygen
c.oxygen
pel Pbulk
(g/cm 33)) (g/cm 3
3
)) (emu/g) (emu/g) /M,.bulk (wI. %) (wI. %)

C020
20 Ni 80
ao 8.81
881 890 0.99 767
76.7 77.7 099 006 0

C0
Coso Ni 50
50 Niso 8.67 8.90 0.97
097 109.2 111 .55 0.98
098 021 020

CoaoNbo
C080 Nbo 840 8.90
890 094 138
138. 1 141 .7 098 042 0.62
062

Feo.
Feo 13 [COso
13 [C0 50 Ni 50
so Jo 87
a7 7.39
739 >8.5
>85 <087 107 125 a o 85 a 17
1.7 35

Feo 13 [Co
FeD 13 ao Ni
[C080 Nbo Jo 87
20 Jo. a7 736
7.36 >8.5
>85 <087 084 1.6

(1) bulk value is not available in the literature,


literature. it has been calculated on the assumption that the
saturation magnetization iinearly depends on iron content.

7. 3. 4. 2
7.3.4. Shell Characterization

The thickness and the overall density of the shell can be inferred as well from
the experimental data (Figs. 7.7 - 7.9) according to Eqs. (7. 2b) and
(7. 2c). The thickness of the superficial layer is about 2 nm in the Cox lOO - x
Ni 1DD
CoxNi
while it is larger in FeCoNi powders. This layer is twice as thick for
FeD. 13 [C0 5D Ni 5D ]D 87 than for the C0
Feo13[CosoNiso]087 5D Ni so
Coso 5D particles for instance (Table 7.2).
Such values are consistent with the thickness of the oxide layer of Fe
(Gangopadhyay et aI., 1992a; Tamura and Hayashi, 1984; Shinjo et al. ,
1980), Ni (Haneda, 1987) or FeCoNi (li
(Li and Takahashi, 2000) particles
obtained by different physical routes. They are also consistent with the
thickness of the oxide layer evidenced by TEM over polyol-made
pol yoI-made nickel
particles (lalla
(Lalla et al., 2000). Nevertheless, the overall density of the
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 241

superficial layer of polyol made particles appear always markedly lower than
those expected for pure metal oxides (Toneguzzo et al. , 2000). Typically, in
the Cox Ni
NilQo-
1oo - x system, the calculated density of the impurities layer slightly
varies around 4 g/cm 3 for different metal compositions (Table 7. 2). The
nature of the superficial layer has been inferred from thermal analysis
associated with thermo-programmed desorption (TPD) , gas phase
chromatography (GPC) and thermomagnetic measurements. The results
(Toneguzzo et al. , 2000) are consistent with a superficial layer formed mainly
by an inorganic moiety (metal oxides and in a low extent adsorbed water) and
to a lower extent by a metallo-organic one (the metal alkoxide hydroxide
observed as an intermediate phase). The lower layer density as compared to
the metal oxides one is therefore well explained by the presence of this less
dense metallo-organic moiety.

Table 7.2 Density p, thickness e, of the outer shell for different compositions;
compositions: the values
are inferred from Eqs. (7. 2b)and (7.
C7 2c)
P, (g/cm 3 )
p,(g/cm e (nm)

Co 20 Ni so
CozoNi 4.0 2.6
Coso Ni so 3.7 1.8
Coso Nbo
CosoNizo 4.2 1.6
Feo.
Feo 13 [COso Niso Jo S7 33 3.2

It must be pointed out that this superficial layer is in most cases thick
enough to prevent the catastrophic oxidation of metal particles under air
exposure. However, such a pyrophoricity phenomenon occurs for the finest
powders (d m <25 nm) due to their high specific surface area (A ss >40 m2 //g).
g) .
this superficial layer is not thick enough to prevent completely
Nevertheless, tnis
the sintering of the particles during the thermal treatment required to lower
non-metallic impurities content (ct.(cf. section 7.3.3.3),
7.3. 3. 3), nor to electrically
insulate the particles in order to obtain a non-conducting material by
mechanically compacting them in order to investigate their dynamic magnetic
properties (cf. section 7. 4. 2. 1). This has been achieved by coating the metal
particles with a thin dielectric layer of hydrated manganese( W)N) oxide through
a soft chemical treatment in solution (Kocon et ai.,aI., 1994; Toneguzzo et aI.,
al. ,
2000). The coating of the particles appears by TEM quite homogeneous with
an average thickness of about 10 nm. The main advantage of this method is to
provide, by compacting the powder, a composite having a high volume
concentration of magnetic inclusions without any electric percolation. The final
characteristics (0 and C content, density) of the superficially treated and
annealed particles of different composition are given (Tables 7.3 and 7.4) 7. 4)
with respect to the characteristics of as-made powders. Furthermore, it must
be stressed that the morphological characteristics of the powders are retained
after such an annealing
anneal ing treatment (Fig. 7. 10).
242 Philippe Toneguzzo et al.

Table 7.3 Oxygen and carbon contents of Cox Ni IOO - x and Feo
NilQo- Feo. 13 x Jo 87 powders for
Ni ,oo - xJo.
13 [Cox NilQo-
different compositions: (a) as-made powders;
powders: (b) superficially treated powders;
(c) superficially treated and annealed powders.
Composition (a) (b)
( b) (c)
(wt.
(wI. %) 0 C 0 C 0 C
Co20 Ni 80
C0 0.88 0.27 3.8 031
0.31 23
2.3 0.05
005
C0 3s
35 Ni 6s
65 1.0 0.36 3.8
38 0.40 22.11 0.07
C0 50 Ni 50
Coso so 1.1 0.36
036 4.1 036 26
2.6 0.08
008
C06s
65 Ni 35
3s 1.2 048 4.7 0.54 2.3
23 o 13
Feo 13
13 [C0
[Co20 Ni 80 Jo 87 4.4 1.6 7.3 1.5 4 034
Feo.
Feo 13 [C0 3S Ni 6s
[Co 35 65 Jo 87 8.4 1.4 3.7
3 7 o 15
Feo. 13 [C0
FeO.13 50 Ni50
[Coso Jo. 87
Niso Jo 8.2 1.7
.7 11.6 4.3
43
Feo. 13 [C0
FeO.13 [Co80 Ni 20 Jo 87 3.5 1.5
.5 6 .4
1.4 32 0.25
025

Table 7.4 Density of Cox Ni ,00


lQo - x and FeO.13 [Cox Ni,oo-
NilQo- x Jo 87 powders for different
compositions (a) as-made powders; (b) superficially treated powders: powders;
(c) superficially treated and annealed powders.
pp(g/cm 33 )
(PCe) - pIa)
(PIc) P,., )) / pIa)
p,.,
Composition dm(nm)
(a) (b) (c) (% )

C0
Co20 Ni 80 315 8.55
855 7.72 858 0.4
C0 50 Ni 50
Coso so 190 41
88.41 7.56 846 06
Feo 13 [C020
Feo. [Co20 Ni80 Jo 87
Ni 80 Jo. 225 7.14 6.57 7.76
776 87
8.7
Feo
Feo. 13 [C0 50 Ni
[Coso 50 J087
Niso JO.87 205 7.06 6.00
600 8.26
826 17

7. 3. 5 Static Magnetic Properties


The polyol process, which allows the synthesis of spherical and monodisperse
ferromagnetic particles with a mean size varying over 3 orders of magnitude,
offers the opportunity to study their static magnetic properties in the size range
where the change from a multi-domain to a single domain configuration is
expected. According to the core-shell structure previously described, the
relevant diameter to consider in this study is the diameter of the magnetic
20-200
metallic core. Indeed, in the 20 - 200 nm size range where the shell thickness
estimat~d to a few nanometers, the difference between the particles
has been estimated
and the core diameter can be neglected. For the finest particles, since the
shell thickness is unknown, the overall particle diameter has been considered
as well even if it leads to systematically overestimate the magnetic particles
diameter.

7.3.5.1 Static Magnetic Properties Versus Particles Size


Saturation magnetization, remanence and coercivity were inferred from room
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 243

(a)

(b)

Figure 7. 10 TEM image of FeoFeo. 13 [COso so Jo


[C0 50 Ni 50 87 particles prepared by the polyol
J087

process through heterogeneous nucleation using Pt as nucleating agent: (a) after


superficial treatment ( d m = 215 nm.
nm, a = 20 nm); (b) after superficial treatment and
annealing under argon (d mm == 200 nm.
nm, a = = 20 nm).
nm) .

temperature hysteresis loop measurements on powders compacted without


non-magnetic binder into a rigid assembly. In such experimental conditions,
conditions. it
is clear that the interactions between particles influence these magnetic
properties. The variation of coercivity He and remanence to saturation ratio
M,/ M s with the mean diameter d m of the polyol made particles is exemplified
in Fig. 7. 11 for the Coso
C080 Ni zo composition. Two size ranges separated by a
Nizo
critical size are clearly evidenced. In the upper size range, range. He and M,/
M,/MM ss
increase linearly versus the reverse of the mean diameter. In the lower size
range, He and M,/ M ss sharply decrease as d m
range. m decreases and reach down zero
Furthermore. the critical size at 300 K Iies
below 7nm. Furthermore, lies in the 20 - 30 nm and
30 - 40 nm ranges for Coso
C0 80 Nizo and for C0
Coso
50 Ni so
50 respectively. For this size.
size,
244 Philippe Toneguzzo et al.

Mrr // M
M Mss reaches the value of 0.45 while He tends to 575 Oe and 370 Oe for
C080 NiNizo
20 and C0 50 Ni 50 compositions respectively.

600

EJ 0
He
0.50

0 o M,JMs

400

i .
i 0.25 :i::{':{~'C

200

0
o.

0
,.
/no
!'no

o o
I 10 100 \000
Rm (nm)

Figure 7.11 Coercivity He (e) and remanence to saturation ratio M,/M, (D) of CosoNi,o
CoaoNbo
( T = 300 K).
particles vs. mean diameter d m (T

This size dependence of coercivity and remanence to saturation ratio and


the occurrence of a critical size have previously been reported in the literature
for other ferromagnetic particles prepared by physical or chemical routes
(Uehori et al., 1978; Yiping et al., 1992; Herzer, 1992; Gong et al. ,
1991) .
The critical size, which is therefore characterized by the highest
coercivity and remanence to saturation ratio, corresponds to the size below
which magnetization reversal of the spins within the single-domain particles
occurs by coherent (or synchronous) rotation. The experimental critical size
can be compared with the theoretical value de inferred from the following
equation (Aharoni, 1996).

de = 2. 034 v75
JC
M (7.3)
(7.3)
s

where C is the exchange constant. According to this equation and assuming


the exchange constant C = 2. 10- 6 erg/
C=2. cm, de is equal to 24 and 30 nm for the
erg/cm,
C080 Ni
Nbo20 and C0 50 Ni 50 compositions respectively. The experimental values are

therefore consistent with these theoretical ones. Furthermore, the highest


experimental M,/Mr/M (i. e. , 0.45) is not so far from the value expected
M ss ratio (i.e.,
for a random distribution of single domain particles (i. e. , O. 50) .
Above this critical size, in the size range investigated (below 100 nm), nm) ,
the magnetization reversal does not occur by coherent rotation but through
inhomogeneous reversal modes such as curling. curl ing.
Monodisperse Ferromagnetic Metal Particles: Synthesis by...
by. . . 245

For the finest particles in the lower size range, a superparamagnetic


behavior is observed above a blocking temperature T b which depends on both
the particle's size and composition. Blocking temperature T b is inferred
experimentally from zero field cooled (ZFC) thermomagnetic measurements
(Fig.7. 12). Below that temperature, coercivity of the nanoparticles increases
when the temperature decreases. For non-interacting single domain particles
with a uniaxial anisotropy, coercivity has been described by:

H =H
H =
c
H e,O
c.O
(1 _ J25k
KV T)
)25k B
T) (7.4)

where k B is the Boltzmann constant, K the anisotropy constant and V the


particle's volume. Such a T 1' //2 variation expected from this equation has been
experimentally observed (Fig. 7. 13).
10.0.----------,-------,------,
10.0,---------,-------,--------,

FC
x ZFC
7.5 f---~-__I----_I_----_1
I---~--I-----I-------l

~E 5.0 ~-----t-~:;_--_I_----__1
t--1i~-------j-__......~--__t----_i
~
~

2.5 1
t --
----
----
---- t
---- j
----
_ -I
-__-
_-t-
---_
--__1
i

o 100 200 300


T(K)

Figure 7. 12
7.12 Susceptibility vs.
VS. temperature in FC/ZFC mode for Coso Ni so particlesC
particles(ddm =
3 nm).

The experimental linear variation of the static magnetic properties versus


the reverse of the mean diameter of the polyol-made particles in this size
range has been early observed in the literature for other ferromagnetic
compounds (Kneller, 1969; Kobayashi et al. , 1995).
7.3.5.2 Static Magnetic Properties Versus Particles Composition
The variation of coercive field for different compositions in the Cox Ni,oo-
Ni 1oo - x
system and for an almost constant mean diameter in the upper size range
(d m
m =220nm) is exemplified in Fig. 7.14. He is minimum for Co20 Ni so
composition and then it linearly increases with the Co content up to pure Co.
246 Philippe Toneguzzo et al.

600 .-----~--~---~---~
r----~--~---~---~

4001---- - 1 - - - - 'I,->-----l-------j
400 \------\-----~I__--__I----~


200 \-----\------II__----4"-._+_
1-----I--------j>---~.'\__1


o 2 4 6 8
rT I2
l2
(K I2 )

Figure 7. 13
,
r I ,22 for C080 Ni 20 particles (d
Cd m = 4 nm).
Coercivity vs. T nm) .

200

<)
0

150
A ~

<)
0
<)
0 <)
0
<)
0
0e
<)
0
<)
0

50 1--.
....6. ....6. 1--._--

o<) CO,Ni IOO _x


... Feo
.6. Feo ulCo,Ni 100-x]
dCo,Ni 100-xlo0 87

I II
o 25 50 75 100
x
Figure 7.14 Coercivity vs. Co content x in CO,Ni IOO - ' and Feo
CO,NilQo-, 13 [Co,Ni IOO - ' Jo
[Co,NiIQO-' 87

sizeC d m = 220 nm) .


systems for particles with an almost constant size(

Iron based particles always present lower He than the CoNi particles whatever
the relative Co/Ni ratio. It is noteworthy that the lowest coercive field for the
C0 20 Ni ao composition can be related to the minimum of magnetocrystalline
Co
anisotropy observed for this composition on Cox Ni lOolOO - x bulk alloys (Landolt-
Bornstein, 1982).
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 247

Furthermore, for the finest granularities in the lower size range, the
blocking temperature T b inferred from Zero-field CoolingCZFC)
Cooling(ZFC) measurements
varies with composition in the Cox Ni Ni,oo-
100 - x
x system for an almost constant
(Table 7.5).
nanoscale particles size CTable 7. 5 ) . T b b increases when Co content
increases.

Table 7.5 Blocking temperature T bb of nanoscaled CoNi particles.


Composition Coso Ni,o
Nbo C0 G5 Ni 35
CO65 C0 50 Ni 50
4.1
4. 2.8
28 3.1
145 115 45

7. 3. S.
5. 3 Calculations of the Magnetocrystalline Anisotropy Constants
The magnetocrystalline anisotropy constants can be estimated from dynamic
and static magnetic properties for submicrometer and nanometer sized
7. 4. 2. 5, the values
particles respectively. Thus, as it will be shown in section 7.4.2.5,
K I are found close to the bulk ones for Cox Ni loo
of the anisotropy constant K, lOo - xx
Cd mm = 200 nm). They increase with x. This behavior is consistent
particles (d
with the increase of the relative proportion of the anisotropic hcp phase.
For particles in the 20 - 30 nm size range, by using the Stoner and
Wohlfart model, i. e. , neglecting particles interactions and assuming a random
distribution of single domain particles with cubic anisotropy whose reversal
mechanism is coherent, the anisotropy constant is roughly estimated by

K - Hc.maxMs
(7.5)
,I -- 0.64

where H c.max is the coercivity at the critical size. In this size range, for
Coso Ni 20
CosoNi zo and C0
Coso50 Ni 5o K I is found equal to 9 x 10
so , K, lOs5 and 5.5 lOs5 erg/
5. 5 x 10 cm 3
erg/cm
respectively, i. e. , sl ightly larger than those calculated in the upper size range.
slightly
Finally, for the finest particles (CosoNi
CCosoNi 20 =
zo ,' d m =44 nm), the K,
K I values are
I
determined either from the slope of H cc vs T /2 curves using Eq. (7.4)
T'/2curves (7. 4) or from
the blocking temperature T busing

K, =
K
I
= 25k s T b (7.6)
V
V

where V is the particle's mean volume, both equations being inferred


assuming an uniaxial anisotropy. The K K,1 values are found equal to 6 x 10 6 and
1. 3 x 10 7 erg/ cm 3 through these two methods respectively. Thus it can be
seen that the anisotropy constant increases steadily as the particle's size
decreases for the Coso Ni zo
20 composition. The high values observed for the finest
particles can be attributed to the increasing influence of the surface
anisotropy.
The inferred values are in fact related to an effective anisotropy constant
that accounts for the magnetocrystalline anisotropy and the surface anisotropy.
248 Philippe Toneguzzo et al.

The following equation (B0dker


(BQldker et al. , 1994) :

6
Ken = KV+(jK
KeH=Kv+(jK s (77)
(7.7)

where K v and K s are the magnetocrystall


magnetocrystalline
ine and surface anisotropy
contributions respectively, has been widely used to take into account the
increasing influence of surface anisotropy as the particles size decreases.

7. 4 Magnetic Resonance in Fine Particles

Recently several works have been devoted to the magnetic resonance, both
from a theoretical and experimental point of view, in very fine particles,
namely with a mean diameter lower than 1 IJm. Two contributions that are
neglected for bulk materials appear significant when the ferromagnetic particle
size is reduced: the exchange field and the surface effect. In the
submicrometer and nanometer size ranges these two contributions lead to a
size dependence of the resonance frequencies. Moreover, the exchange field
induces that several non-uniform modes can be excited by a uniform r. f. field
in fine particles. Theoretical calculations were done within a continuum model
(Aharoni, 1991, 1997; Shilov et al., 1999a, 1999b) or with a discrete
treatment as well (Mercier et ai., al., 2000). In section 7.4. 1, the main
theoretical results for a spherical particle will be recalled.
Magnetic resonance experiments are performed on classical FMR
spectrometer (Gazeau et al., 1998; Bakuzis et al., 1999; Respaud et al. ,
1999), on microribbon microwave transmission devices (Goglio et al. , 1999;
Encinas-Oropesa et al.,
ai., 2001) or on coaxial Iine line associated to a network
analyzer (Viau et al. , 1997; Toneguzzo et al. , 1998). The two lalter
latter devices
present the interest of large frequency range measurements, from 100 MHz to
18 GHz or 40 GHz. In section 7.4.2, results on magnetic resonance
experiments performed on Co-Ni and Fe-Co-Ni particles prepared by the polyol
process will be presented.

7.4.
7. 4. 1 Theoretical Background

In the general theory of ferromagnetic resonance, the motion of the magnetic


moments within a ferromagnetic particle is formulated by the Landau-Lifshitz
equation:
um = y[m 1\ Hen + am 1\ (m 1\ H on )] (7.8)
Jt
Monodisperse Ferromagnetic Metal Particles: Synthesis by...
by. . . 249

where m is a unit vector parallel to the magnetization vector m = M/ M s ' is


(the gyromagnetic ratio is, a, the damping parameter and H eH is.
Heft The
effective magnetic field in the medium. H eH is the sum of several contributions:
Heft

H
Heft
eH
=Q V 2 m - _1_ <Jw
'\12 dW a +H (7.9)
Ms
M M
Ms <Jm
dm

where M ss is the saturation magnetization, C = = 2A is the exchange constant,


W a is the anisotropy energy density and the magnetic field H = H o + H m m is
composed of the external applied field, H o , and the dipole field created by the
volume and surface charge of the magnetization distribution, H m .
From Eq. (7.8) and Eq. (7.9), several simplifications are generally
made. First of all, the damping is generally ignored. Note that damping was
included by Voltairas et al. in calculation of magnetization configuration in
small sphere (Voltairas and massalas, 1993). Second, the magnetostatic
energy is neglected; that means that the magnetostatic interactions between
spins leading to terms other than the demagnetizing field are not taken into
account in H m . This last assumption is valid only for a very small sphere. For
micrometer sized particles, mode mixing was calculated by Voltairas et al.
(Voltairas et a!.
al. , 2000a,2000b).
The following equations are established by neglecting damping and
magnetostatic energy and assuming a resonant process with an oscillating part
of the magnetization mx = my cc exp ( - iwt) where w is the resonance
frequency, the z axis being collinear
coli inear both with the anisotropy field and the
external field. It means that a uniform orientation of the effective magnetic field
is assumed within each grain, this grain being a single-domain or containing
only a few magnetic domains whose common crystalline orientation defines the
crystalline anisotropy easy axis. Thus, no external field is required to
magnetically orient the sample and in such a case which is close to the
experimental one, the effective field is just the anisotropy field with addition of
the demagnetizing field due to the magnetized parts. For the whole grain set,
this approach means that all grains share a common crystalline and magnetic
orientation. As a matter of fact, such a common orientation between grains is
obviously imposed by dipolar interactions which define the demagnetizing field
in the sample. This common orientation has been observed in numerous
numerical computations of the magnetic and structural nematic-like order of
grain chains when magnetic spheres interact via magnetic dipolar interactions
and short-ranged repulsive interactions (Wei and Patey, 1992; Weis and
Levesque, 1993). Thus, this common "nematic" anisotropy axis is well
experimental
justified here from dipolar considerations even in the present experimeotal
situation of magnetic resonance without external field (cf. section 7. 4. 2. 1).
To a first order such an approach leads to the linearized equations:

(7. lOa)
250 Philippe Toneguzzo et al.

(( ,,2_ MsHz)
V2_MsHz)
C m x
~ omy = 0
__~amy=O
YoC ot
at . (7. lOb)
0.10b)

For a spherical particle of either uniaxial or cubic anisotropy provided z is an


easy axis Eq. O.
(7. 10) lead to:
,
=
Hz = H - 41T
4iT M
o 3 s
+ 2K
M
1
(7.11)
0.11)
s

where K 1 is the anisotropy


an isotropy constant.
As far as finite samples are concerned, the boundary conditions imposed
on the magnetization vector at the particle surface are determinant for the
resonance equation.
If there is no surface anisotropy, the boundary conditions are (Aharoni,
1991) :

am
(om
x ) _ (amy)
(om y ) = 0 0.12)
(7. 12)
or
ar r=R or
ar r=R
where R is the radius of the sphere.
Equation (7. 12) assumes free surface spins or in other words that the
spins located at the particle surface are equivalent to those located inside the
bulk.
Under this last assumption, Aharoni solved Eq. (7. 10) by means of
Bessel functions (Aharoni, 1991). It leads to a set of resonance frequencies,
w, given by:
= Q J.1~n
w = ,u~n + H _ 41T M + 2K ,
4iT M (7.13)
0.13)
Y Ms R 2
M 0 3 s Ms
M

where the eigenvalues ,ukn


J.1kn are the roots of the equation:

jn
djn,ukn
( d J.1kn
r/ R ) =0, n~O 0.14)
(7.14)
dr r=R

where jn are the spherical Bessel functions. The first ,uknJ.1kn values have been

calculated by Aharoni (Aharoni, 1997).


The lowest frequency mode of Eq. (7. 13) has the zero eigenvalue. It
corresponds to the well-known uniform mode given first by Kittel (1948). In
the absence of surface anisotropy it does not vary with the particle size. The
following modes are called exchange resonance modes and scale for small
particles as a quadratic function of the particle reciprocal size.
The size dependence described above for both the uniform and non-
uniform exchange modes are no longer valid in case of pinned surface spins,
meanly when a surface anisotropy modifies the boundary conditions.
In the continuum model the uniaxial anisotropy energy density on the
surface has been introduced on the form (Aharoni, 1997; Shilov et al. al.,,
1999a, 1999b):
ws =K s (1 - m;) (7. 15)
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 251

where K s is the surface anisotropy constant.


It modifies the boundary conditions of Eq. (7. 12) into:

am
d + 2 K ssm
m xx +2K m _am +C
= d m yy +2K
2 K ssm = 0, on r = R
m =O,onr=R
ar c
7i( C xx-7i(
dr C Yy
(7.16)

where appears the critical size R e :

_ C (7.17)
Re - '
2Ks
These new boundary conditions change the size dependence of the uniform and
the exchange resonance modes for small particles anti for large particles,
respectively.
( 1) For R / R e 1, Shilov et al. showed that the uniform mode, which
does not involve the exchange energy, becomes non-uniform (Shilov et al. ,
1999a, 1999b). Its precession frequency in the case where both the anisotropy
field and the external field are assumed to be collinear with the z axis, is
given by:
w
W = H _4TT M +2K , + 3K s (7.18)
y 0 3 s Ms MsR'
Thus the surface effect results in an additional size-dependent field which
scales as R- 1
(2) At the opposite end for R / R e
1, the magnetization inside the
particle is quasi uniform and no size dependence is expected.
As far as exchange modes are concerned, the eigenvalues JJkn J.lkn are no
longer constants but become size dependent when the boundary conditions
read as Eq. (7.16)
(7. 16) (Aharoni, 1997).
If R has the same order of magnitude as R e' the size dependent term of
Eq. (7. 13) is modified, it scales as R - f3~ with f3 lower than 2, and can reach
values as low as O. 5.
At the opposite the R- 2 size dependence given by Eq. (7.13) is still
expected for small particle size, meanly for R much smaller than R e .
The main results on the size dependence of the resonance frequencies are
found as well by the discrete model (Mercier et al. , 2000). According to this
model the size dependence of the resonance frequencies is also mainly related
to the surface pinning and to the exchange parameter value.
In each cases the size dependence of the lowest frequency mode is found
weaker than that of the following modes. Moreover, either in absence of
surface pinning or for large particles, the frequency of this first mode varies
very slightly with the particle size and is directly related to the crystalline
anisotropy. The size dependence of the higher modes is found to scale as R- 2
either when the particle size is decreased or when the exchange parameter is
increased (Mercier et al., 2000). These results agree well at least
qualitatively with those inferred from the continuum model.
To conclude this part, it must be stressed that the models presented here
252 Philippe Toneguzzo et al.

deal with resonant effects in independent small spherical particles. Further


works are necessary to account for the magnetic interactions between grains
and to study their possible influence on the resonance frequencies.

7. 4. 2 Magnetic Resonance in Polyol


PolyoI Made Particles
7. 4. 2. 1 Experimental Section
The dynamic magnetic properties of the ferromagnetic particles were
characterized in the 1'00 MHz - 18 GHz frequency range with an APC7 coaxial
Iine associated with a network analyzer HP 8720. With such a method the
microwave properties [E ((w), 11 ( W ) J are measured on non-conducting
W ), J.1
composite materials having high volume concentration of ferromagnetic
particles. Measurements are conducted in the absence of an external magnetic
field.
Composite materials were prepared by randomly dispersing the powqers
powders
in an epoxy resin or an elastomeric matrix with a volume concentration ranging
from 10% to 40 %. The particle concentrations in composite materials were
determined by density and saturation magnetization measurements. For the
finest particles, the difficulty to obtain homogeneous materials by dispersion in
the matrices led us to coat the particles by a chemical treatment (cf. Section
7.3.4.2). The ferromagnetic particle volume concentration in the compacted
samples was determined with accuracy from the mass and the volume of the
sample, the saturation magnetization of the coated powder and from saturation
magnetization and density of the as made powder. This concentration lies
between ca. 25 % for the finest particles and 50 % for the largest ones.
The permitivity, E (w), of the composite materials was found to be low
and constant on the whole frequency range studied. No dielectric losses were
observed (E"(W) = 0) .
(EN ( W ) =0).

7. 4. 2. 2 Intrinsic Permeability

The microwave properties of composite materials are very dependent on the


magnetic particle volume concentration. This influence has already been
described with some success by a homogenization law adapted from the
Bruggeman effective medium theory. For spherical ferromagnetic inclusions
and neglecting the contributions of eddy currents to the permeability (which is
~m), one can write (Berthault et al. ,
justified for mean diameter lower than 2 IJm),
al.,, 1993):
1992; Rousselle et al.
J.1e[2J.1e + I1
l1e[2l1e m C3q - 2)J
J.1m(3q
(7.19)
l1i =
J.1i e C3q - 1) + 11m
J.1e(3q
l1 J.1m

where l1i
J.1i is the intrinsic permeability of the ferromagnetic particles, 11m
J.1m is the
permeability of the matrix (J.1m = 1 in our case) and q is the particle volume
(11m =
fraction.
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 253

2.0
0,,-
----
-----------,r-------------,,-
----
-- -- - - - ,
x 20%
30%
030%

Xx 040%
'\ x 45%
/------_+~x:<-----'x'-x-\--------j
1.5 f-------i--->x'------'x'--x-+-------1

="
=::t 1.0
1.0 I - - - - - - - - - " - h ' f ' - - - - - - i \ ! r - - - - - - - i
::l.

0.5 f------H--t-.#"----------::>''-----N~--------1
1-----e--h9"--------:>''----N~--_____j

1 10 100
Frequency (GHz)
(a)
s,--------r------,--------,
030%
040%
x 45%

6 I-------+H-----+------j

__ 4 I---~--,.,.+-- ---~I1.t_----__I
-::t

g
2 /--,(f--i0~8-_+-----a----_l
g
~x
00
~ x
.Ix<> X
O'-- --'- .l....- ....J

0.1 I 10 100
Frequency (GHz)
(b)

Figure 7. 15 /1': (( w) of composite


(a) Imaginary part of the extrinsic permeability JJ':
materials made up of Coso Ni so particles dispersed in elastomeric matrix at different
volume concentration; (b) imaginary part of the intrinsic permeability JJ;'( w) of Coso Ni so
/1;'(w)
particles calculated from /1; ( w) values at different volume concentration.
JJ: (w)

According to these studies, the intrinsic permeability J.li


J.1i ( w) of the
powders was computed from J.leJ.1e (w) measurements for several composites of
different volume fractions (Fig. 7. 15a). For a given powder (defined by its
composition and particle size) the values J.l;
J.1; were found weakly dependent on
254 Philippe Toneguzzo et al.

the volume fraction, provided that it is higher than 25 % (Fig. 7. 15b).

7.4.2.3
7. 4. 2. 3 Micrometer Sized Particles
The dynamic permeability of Co-Ni particles in the micrometer range presents
a single and broad resonance band in the (0. 1 - 18 GHz) frequency range
(Viau et al.,
aI., 1994, 1995) (Fig. 7.16). 7. 16 ). Moreover, these bands are
asymmetrical with tails toward high frequencies. A shift of the resonance band
with the composition is observed; it can be interpreted qualitatively as the
influence of the magnetocrystalline anisotropy field. It is found that the
maximum of the experimental IJ' /-I': curves, ff"r , and the H a values of CoNi bulk
alloys reported in literature vary in a same way with the Cox Ni 100 - xx
Ni,oo-
composition. The resonance band is shifted toward low frequencies when Ni
content increases in CoNi powders in agreement with decreasing anisotropy
values, and the lowest resonance frequency is observed for the C0 20 Ni ao so
composition which corresponds to a bulk alloy of lowest magnetocrystall ine
magnetocrystalline
anisotropy. However, the experimental resonance frequencies are always
higher than the Kittel's frequency for the uniform mode of a sphere given by
f = ( y /2lT)
/2n) H a' Tthis is not surprising since the particles are far from being
single domain in this size range.
7
7.5. 5 , - - - - - , - - - - - - - , - - - - - . . . . . . ,

j ~"\
oa
ob

0
0
0
~o'x~
of 0 0
oc
xd
o 0 o ,; 00 ~\ "0
5.0 I--------i't--:"---'n--.r--;,-:\f--------j
o ~ xx 0
0 0
0
0
0 0 0
00 x o 0 0
0
0 0
Do
~ x
X o ~
o ~ S
o '6
0
x
x
x
\
<f o \',
2.5
1 00.
x
,
1------v-+--------'~:-t'k-_+----1
~\
I
~ xx ors
\~
o~'" 1
0
0"0 ~ o~ 0

i/ ~
0

j/
oO_~-----'---------'------"'--------'
~.
0

000

0.1 II iO
10 100
Frequency (GHz)

Figure 7. 16 Imaginary
imaginary part of the intrinsic permeability IJI ( w) for micrometer sized
/11 (w)
= ,.
Co-Ni particles: (a) Ni d mm = 1. 4 ~m; Co20 Ni so d mm = 1. 5 >1m;
>1m; (b) C0 ~m; (c) C0 = ,.
so d m = 2 ~m;
50 Ni 50
Coso >1m; =
(d) CosoNbo d m =2 ~m.
>1m.

Actually the particle size lies in a range where both magnetostatic and
exchange interactions are of the same order of magnitude and mixing of the
modes can take place. Voltairas et al. analyzed the dynamic properties of the
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 255

micrometer sized Co-Ni particles in these terms (Voltairas et al., 2000a,


2000b). The exchange constant A (A = C /2) and the 9 factors were
2~g~2. 0).
calculated. The 9 factors were found in the expected range (2. 2;):g;):2.
The exchange constants were found to increase steadily with the Co content in
the Co-Ni composition (from 2 x 10- 7 erg/cm for Ni to 10 x 10- 7 erg/cm for
Co) and to be very close to values determined from thin film measurements.

7.4. 2. 4 Submicrometer and Nanometer Sized Particles


For particles with a mean diameter lower than about 400 nm, several bands
appear in the permeability curves for CoNi (Fig. 7. 17) and FeCoNi particles
(Fig. 7. 18)
(Fig.7. as well, in the frequency range investigated (Viau et ai., 1997;
18)as
al.,, 1997, 1998,
Toneguzzo et al. 2000).
1998,2000).
10 -- a
---- b
.......
--_. c
-_.-
-.--- d
7.5

I""\\
'"
I
\
If I\
I\
5.0
-::t
-:::s..

2.5

1I 10 100
Frequency (GHz)

Figure 7. 17 Imaginary part of the intrinsic permeability 11'


7.17 ,./:; (w)
Ccv) for submicrometer sized
Ca) Ni d m =220 nm;(b)
Co-Ni particles: (a) nm;Cb) C0 ao d m
2o Ni 6o
Co,oNi m =250 nm;(c)
nm;Cc) CosoNi so d mm =250 nm;

(d)
Cd) C0 6o Nbo d m =220 nm.
CoaoNi,o

The accurate control of the particle size allowed by the synthesis process
affords the possibility of following the size dependence of the different
resonance bands for several chemical compositions. For the largest particles,
several resonance bands close to each other are observed. Both the
resonance band frequencies and their relative intensities are found to vary with
the particle mean size. As exemplified in Fig. 7. 19 for C0
COSO Nbo
8o Ni 20 particles, when

the particle size decreases an overall shift of the bands toward high
frequencies is observed and the relative intensity of the lowest frequency
resonance band increases steadily. For the smallest particles, the bands
become less and less close to each other, the number of bands observed in the
range studied O. 1 - 18 GHz decreases and the relative intensities of the
highest frequency resonance bands decrease. Below a mean diameter, which
256 Philippe Toneguzzo et al.

12 ,-------,-~:'\.-.~ - - - - - - r - - - - - ,
12 ,-----,-~,------,------,
:\
::. ..
:' "
,

--3
....... b
9 f-------:-,,\--_+--ei---j('\--:",_-+-_"._,,_._b_--j
9 f------"-;----t----<--/--'r---t--------I

-::I.
-::t
U\
6 f---i---e-----j:-\--k-\J.-------<H-------j
r---i---<----jI-\--fc-\J..------<H-------j

3 f---!----+-:-tt/

o'
0.1 1 10 100
Frequency (GHz)

Figure 7. 18 Influence of iron content upon the imaginary part of the intrinsic
permeability for two powders with an almost constant particles size: (a) C035 Ni 65
m=
dm =244
244 nm; (b) Feo
nm,(b) 13 [C035 Ni 65 Jo 87 d m = 226 nm,
Fe013[C035Ni65]087dm=226 nm.

1
122,,----------
---,,- - - - - -
-,,.-.-
----
----
-- ,
............ 3
.... ..3

- - - b
--c
9 f-------+-+--\----t-------I
9f--------.+-t---+----+------1

=:i
-::t 6f-------+I-i----'-+----If-Ir-It-------I
6f-------+f-7-----'--\---IJ-1i--1'r------1

3 f-------/f.-y-....;-+-';-;/---+-\\\------1
f-------Hi-+-+-----;..LJ--+-'jl\------J

I1 10 100
Frequency (GHz)

Figure 7. 19 Imaginary part of the intrinsic permeability vs.


VS. frequency of C080 Ni zo
20

particles: (a) d m
m =142 nm;(b)
nm,(b) d m
m =82 nm;(c)
nm,(c) d m
m =62 nm,
nm.

is about 50 nm for every composition studied, a single band only remains in


that range (Fig. 7.20).
Despite an overall shift when d m decreases, it is noteworthy that all the
resonance frequencies do not present the same size dependence, a far less
important shift being observed for the first band than for the following ones.
Monodisperse Ferromagnetic Metal Particles: Synthesis by... 257

16
16,--------,----,...-----------r--------,
r-------,----,.;,---------,-------,

12
121------t--/-----'~-_+----_I
r-----f--+-----'~--+----._1

-::I. 8

100

Figure 7.20 Imaginary part of the intrinsic permeability vs. frequency of


Co x Ni lOo
Cox lOo -
- xx nanoparticles: (a) C0 3S
35 Ni 6s m = 50 nrn:
65 d m nrn; (b) Coso
C050 Ni 50 m = 50 nm;
so d m
CoaoNho d m =34
(c) CosoNbo = 34 nm.

7. 4. 2. 5 Size Dependence Analysis


A rough calculation taking C = = 4 X 10- 6 erg/em,
erg/cm, Ms == 103 emu/cm3 ; ' KK I1 = =
5 3
10 erg/cm
erg/ cm , shows that the exchange term of Eq. (7. 13) is very small for a
=
sphere of R = 1 IJm but becomes significant when the sphere radius is
decreased to R = = O. 1 IJm. Actually, it is in this critical size range that
experimentally the broad resonance band of the permeability curves resolved
into several peaks under the influence of exchange energy.
Fig. 7.21 shows the size dependence of the experimental resonance
frequencies in 10g10
IOg10 -Iog 10 representations for several compositions. The power
-log lo
law is always found to be lower than O. 5 for the first mode. It is still weaker if
the size range is restricted to the particle diameter higher than 50 nm. These
observations agree qualitatively with the theoretical results presented in the
previous section.
Thus, the low size dependence of the first mode allowed us to estimate
the anisotropy field H a = = wo/Y from the first resonance frequency Wo of large
submicrometer sized particles (Mercier et al. , 2000) at the opposite of the
case of micrometer sized particles for which the exchange modes are not
resolved (cf. previous section). The K, K I values of Cox Ni 100 - x particles have
CoxNi,oo-x
been calculated from the resonance frequency Wo for the largest particles
( dm =
(d = 200 nm) assuming that K 2 can be neglected. K 1I increases with
increasing cobalt content x (Fig. 7.22).
7. 22). That result agrees with the expected
increase of anisotropy with cobalt concentration of such samples because of
258 Philippe Toneguzzo et al.

the strong crystalline anisotropy of cobalt. Nevertheless, the variation does


not exceed one order of magnitude; it is not surprising since hcp phase is
observed only as a minor phase for cobalt rich composition. A very similar
variation was found in previous works for fcc cobalt-nickel bulk alloys and our
experimental value for the Coso so composition ( - 2. 35 x 10
C0 50 Ni 50 erg/ cm 3 ) is
lOs5 erg/cm
close to the bulk one which lies between - 1. 5 x X 10
5
2. 75 x lO
lOs and - 2.75 10 S5 erg/ cm 3
erg/cm
(Landolt-Bornstein, 1982) .
100
100 c - - - - - - - , - - - - - - - - - - - - - - - ,
+ fo
p=-O.72 o f,
b. f2

x f)
I 0 f-----~c__--"k--~----_____1
10

"N
N
::c
:c
o0
~
1--
+

O. I1 '--_..L----'--'---JL.L-LLLL-_-'---.l---L-l.-L-LI.-LJ
'-----_.l....--'--L-.l-LJ-LL'-----_L-....L---'-JLLJ--LW
10 100 1000
dm (nm)
(a)
1 00c-------,----------,
100,------------,----------------,

p=-1.0
p= -1.0

10
10~-----~d----------1
l:----------"d-----------j
p=-0.95

N
"N
::c
:c
o
~
<:::, p=-0.49+
p=-0.49

O. I '--_..L----'---L-l.-L-L.Ll..L-_--L--L---'----L.L...L.Ll.J
0.1 '-----_~....L__-'-LLL.Ll_":_:_-....L------'-----'----L.L...L.Ll.J
10 100 1000
ddm(mn)
m (nm)
(b)
7. 21 Resonance frequencies of microwave permeability vs. mean diameter in
Figure 7.21
a 10g10
log,o -log
-Iog,o10 representation: (a) Coso
CO ao Nbo particles; (b) C0
Coso50 Ni 50
so particles;
(c) CO
C020zo Ni ao
so particles.
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 259

10 .......-----~-r------____,
10,-----------,-------------,
p=-O.92
p=-0.92

p=-0.88

N
:c
o p=-0.16
~

L--~---'--------L-----L---'---.LLLLJ_ _-'---L-L-LJ-LJLLJ
O. 1I L-_--'-----'---'----'---'--'--'-LL-_--'-------'----'--LLLLl.J
10 100 1000
ddm(nm)
m (nm)

(c)
(e)

Figure 7. 21 Continued from page 257.

I

~

E
~
-2
i:!l
'"
I'
o
~ -3
.
;,z

-4

-5
o 25 50 75 100
eontent,xx
Co content,

Figure 7.22 Anisotropy constant K 1 of Cox NilQO-


Ni IOO - x particles vs. cobalt content
"'0
x inferred from the first resonance frequency wo of large particles (see text).

It is noteworthy that the R -1 dependence expected for this mode at very


small sizes was experimentally observed in several studies dealing with
resonance of iron oxides particles in the nanometer size range (Gazeau et al. ,
1998; Bakuzis et al. , 1999).
260 Philippe Toneguzzo et al.

For the higher modes the power law is found close to 1 in the size range
studied. This is related to the effect of the surface anisotropy on the size
dependence of the exchange modes. Following Aharoni it means that the
particle sizes are in the same order than the critical size RR e' This one was
estimated at ca. 50 nm. That means that the R R-- 22 dependence is expected for
particle diameter below 20 nm (Aharoni,
(Aharoni. 1997). Actually an increase of the
experimental power law for the smallest particles probably occurs since a
lie in the O. 1- 18 GHz range for
second resonance band should be expected to Iie
the 25 - 50 nm sized particles if the R - 1 dependence was supposed to be
effective over the whole size range studied.

7. 4. 2. 6
7.4. Control of the Microwave Permeability
Besides theoretical results on spin dynamics in fine particles,
particles. this study shows
also that it is possible to control the main features of microwave permeability
of a composite material, namely the working frequency, the band width and
the permeability level, through the control of the size and composition of the
ferromagnetic particles.
First of all,
all. the full width at half-height varies from several GHz for
micrometer sized particles to a few GHz for nanoparticles (Figs. 7. 16 and
7.20) .
The control of the resonance frequency with the particle size has been
described extensively in the previous section. Moreover, for a given particle's
size, the permeability depends on the particle composition. For micrometer
and submicrometer sized cobalt-nickel spherical particles, the resonance
frequencies are shifted towards high frequencies when the Co/Ni ratio
increases. This has been interpreted as the influence of the magnetocrystalline
anisotropy of the particles. Whatever the size range. range, it is then possible to
adjust the resonance frequency by acting upon the chemical composition. This
has been exemplified for Co-Ni composition. Furthermore,
Furthermore. is the effect of a
small amount of iron in the composition observed through the comparison
between the CoxNi,oo-x Feol3[CoxNilOo-x]o87 systems is spectacular.
Co xNi 100 - x and the FeoI3[CoxNilOo-x]o87
The resonance bands of the iron-based particles appear at lower frequencies
than those of the iron-free particles with the same Co/Ni ratio and a similar
particle size (Fig. 7.
7.18)18) (Toneguzzo et al. , 2000).
Moreover, the comparison of Co-Ni and Fe-Co-Ni permeability curves
showed that the maxima of permeability reached with iron-based particles are
always higher than those reached with cobalt-nickel particles. These higher
permeability levels are obtained despite the fact that the iron-based particles
present a lower crystallinity and a higher impurity content. Moreover, mild
thermal treatments also allowed us to increase the permeability levels. In both
cases it was interpreted by the higher saturation magnetization of the
ferromagnetic core of the particle's (Toneguzzo
(ToneguzZQ et al.
al ., 2000).
Due to their remarkable morphological characteristics and to the accurate
control of their particle size,
size. the polymetall ic magnetic powders made by the
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 261

polyol process appeared as very convenient materials in order to study the


influence of the particle size upon the microwave permeability. Three different
behaviors have been evidenced in the size range where one changes from a
multi-domain to a single-domain configuration. Moreover it appears that it will
be possible by controlling the size and the composition of the magnetic
inclusions to get granular materials with optimized absorption characteristics in
the microwave range.

7. 5 Conclusions and Perspectives

It has been shown during the last decade that chemical precipitation,
precipitation. which is
a low-cost,
low-cost. tasily performed and versatile synthesis method.
method, has potential for
providing monodisperse ferromagnetic metal particles
particles. Controlled nucleation
and growth along with a steric stabilization provide efficient routes to make
nanoparticles with well-defined characteristics. These particulate materials
offer the opportunity to study magnetic properties related to surface or quantum
size effects. Processes to induce the self-organization of such monodisperse
ferromagnetic particles have been developed as well. The first results
concerning the collective properties of long-range ordered aSSemblies
assemblies of those
particles are now available (Legrand et al. ,2000; Petit and Pileni,Pileni. 1997;
Petit et al., 1998; Quid Ely et al.,
al.. 1999). Nevertheless, the study of size
dependent resonance effects in ferromagnetic particles in the presence of an
external radio-frequency field requires the preparation of monodisperse
particles lying in a more extended range than the nanometer one. Thus. Thus,
according to theoretical modes,
modes. size dependent resonance effects are
expected in the submicrometer range where exchange interactions become
predominant over magnetostatic forces. Moreover, surface anisotropy effects,
effects.
which depend on the particles size as well, are expected for particles with a
diameter lower than 100 nm. Model materials to study such size-dependent
resonance effects are now available by means of the polyol process that allows
the synthesis of spherical, monodisperse ferromagnetic metal particles of
various compositions with a mean diameter varying over 3 orders of magnitude
in the required scale range. The large size and composition ranges available
by this synthesis process provide an extended set of resonance spectra which
allow us to bring experimental evidence of exchange resonance modes, to
follow accurately the dependence of their resonance frequencies with particles
size and composition and to compare with theoretical results. The respective
influence of the particle's size, magnetocrystalline anisotropy and surface
anisotropy has been evidenced. Experimental data agree qualitatively
qual itatively with
theoretical calculations based either on a continuum model or a discrete
treatment of the resonant effect in independent fine spherical particles.
262 Philippe Toneguzzo et al.

A deeper comparison between theoretical and experimental data would


require us to take into account the magnetic interactions between particles and
their possible influence on the resonance frequencies through multiple
scattering within the set of particles. Thus, both a more sophisticated
theoretical approach and a certain control of the organization of the magnetic
particles within the composite material in this size range will be needed in the
future. Nevertheless, it already appears that it would be possible by
controlling the size and the composition of the magnetic inclusions to get
granular materials with optimized absorption characteristics in the microwave
range.

References
Aharoni, A. J. Appl. Phys. 69: 7762 (1991)
Aharon
Aharoni,i, A. Introduction to the Theory of Ferromagnetism. Clarendon Press,
Oxford (1996)
Aharoni, A. J. Appl. Phys. 81: 830 (1997)
Antonsen, D. H. and J. H. Tundermann. Metals Handbook, 9th edn. Vol. 7.
ASM Metals Park, OH, p. 134( 1984)
Bakuzis, A. F., P. C. Morais and F. Pelegrini. J. Appl. Phys. 85: 7480
(1999)
Berkowitz, A. E. , W. J. Schuele and P. J. Flanders. J. Appl. Phys. 39: 1261
((1968)
1968)
Berthault, A.A.,, D. Rousselle and G. Zerah. J. Magn. Magn. Mater. 112: 477
(1992)
B0dker F. , S. M0rup and S. Linderoth. Phys. Rev. Lett. 72: 292 (1994)
Bonneman, H. , W. Brijoux and T. Joussen. Angew. Chem. Int. Ed. Engl.
29: 273 (1990)
Cao, X., Yu. Koltypin, G. Kataby, R. Prozorov and A. Gedanken.
J. Mater. Res. 10: 2952 (1995)
Chen, J. P., C. M. Sorensen, K. J. Klabunde and G. C. Hadjipanayis.
Hadj ipanayis.
J. Appl. Phys. 76: 6316 (1994)
Choukroun, R. , D. de Caro, S. Mateo, C. Amiens, B. Chaudret, E. Snoeck
and M. Respaud. New J. Chem. 22: 1295 (1998)
de Caro, D.,D. , T. Ould Ely, A. Mari, B. Chaudret. Respaud.
Chaudret, E. Snoeck, M. Respaud,
J.M. Broto and A. Fert. Chem. Mater. 8: 1987 (1996)
Dinega, D.P.
D. P. and M.G.
M. G. Bawendi. Angew. Chem. Int. Ed. 38: 1788 (1998)
Dong, X. X.L., Z.D.
L. , Z. D. Zhang, X.
X.G.
G. Zhao, Y.Y.C. S.R.
C. Chuang, S. W.M.
R. Jin and W. M.
Sun. J. Mater. Res. 14: 398 (1999)
Du. Y. -W., M.-X.
Du, Y.-W., M. -X. Xu, J. Wu, Y. -B. Shi, H.-X.
Y.-B. H. -X. Lu and R.-H.
R. -H. Xue.
J. Appl. Phys. 70: 5903 (1991)
Encinas-Oropesa A., M. Demand, L. Piraux, U. Ebels and I. Huynen.
J. Appl. Phys. 89: 6704 (2001)
Fievet. F.,
Fievet, F . J.-P. Lagier and M. Figlarz. M.R.S. Bull. 14: 29 (1989a)
Monodisperse Ferromagnetic Metal Particles: Synthesis by. . . 263

Fievet, F.,
F. , J.-P.
J. -Po Lagier, B. Blin, B. Beaudouin and M. Figlarz. Solid State
lonics 32/33: 198 (1989b)
Fievet, F. In: T. Sugimoto ed. Fine particles: Synthesis, Characterization
and Mechanism of Growth. Marcel Dekker, New York, p.460 (2000)
Gangopadhyay, S., G. C. Hadj ipanayis, S. I. Shah, C. M. Sorensen,
Hadjipanayis,
K.
K.J.J. Klabunde, V. Papaefthymiou and A. Kostikas. J. Appl. Phys.70:
5888 (1991)
Gangopadhyay, S., G. C. Hadjipanayis, B. Dale, C. M. Sorensen,
K.J. Klabunde, V. Papaefthymiou and A. Kostikas. Phys. Rev. B 45:
9778 (1992a)
Gangopadhyay, S., G.C. Hadjipanayis, C.M. SorensenandK.J. Klabunde.
IEEE Trans. Magn. 28. 3174 (1992b)
Gangopadhyay, S., S. , G. C. Hadjipanayis, C.M.
G.C. C. M. SorensenandK.J.
Sorensen and K. J. Klabunde.
IEEE Trans. Magn. 29: 2602 (1993)
Gangopadhyay, S., S. , Y. Yang, G.C.
G. C. Hadjipanayis, V. Papaefthymiou, C. M.
C.M.
Sorensen and K. J. Klabunde. J. Appl. Phys. 76: 6319 (1994)
K.J.
Gazeau, F., J.-C. Bacri, F. Gendron, R. Perzynski, Yu L. Raikher, V. V.1.I.
Stepanov and E. Dubois. J. Magn. Magn. Mater. 186: 175 (1998)
Goglio, G., S. Pignard, A. Radulescu, L. Piraux, I. Huynen, D.
Vanhoenacker, A. V. Vorst. Appl. Phys. Lett. 75: 1769 (1999)
D. V. and E. Matijevic. New J. Chem. 22: 1203 (1998)
Goia, D.V.
Gong, W.,
W. , H. Li, Z. Zhao and J. Chen. J. Appl. Phys. 69: 5119 (1991)
Grinstaff, M. W.W.,, M. B. Salamon and K. S. Suslick. Phys. Rev. B 48: 269
(1993)
( 1993)
Hadjipanayis, G.C.,
G. C. , S. Gangopadhyay, L. Yiping, C.M.C. M. SorensenandK.J.
Sorensen and K. J.
Klabunde. In: G. C. Hadjipanayis and G. A. Prinz eds. Science and
Technology of Nanostructured Magnetic Materials. Plenum Press, New
York, p. 497( 1991)
Haneda, K. Can. J. Phys. 65: 1233 (1987)
Haneda, K. and A. H. Morrish. IEEE Trans. Magn. 25. 2597 (1989)
A.H.
Hansen, M. , R. P. Hell iot and K. Anderkop. Constitution of Binary Alloys,
2nd edn. McGraw Hill, New York (1985) and Refs therein
2ndedn.
Herzer, G. J. Magn. Magn. Mater. 112: 258 (1992)
Hisano, S. and K. Saito. J. Magn. Magn. Mater. 190: 371 (1998)
Jezequel, D. , J. Guenot, N. Jouini and F. Fievet. J. Mater. Res. 10: 77
(1995)
Kitakami, 0., H. Sato, Y. Shimada, F. Sato and M. Tanaka. Phys. Rev. B
56: 13849
13 849 (1997)
Kneller, E.F.
E. F. and F.E.
F. E. Luborski. J. Appl. Phys. 34: 656 (1963)
Kneller, E. F. In: A. E. Berkowitz and E. F. Kneller eds. Magnetism and
Metallurgy, vol 1, Chap. vm. VlD. Academ
Academicic Press p. 226 ( 1969)
Kobayashi, K., R. Skomski and J. M. D. Coey. J. Alloys Comp. Compo 222.
1 (1995)
Kocon, L. , H. Lacampagne, P. Latare. In Proc. Powder Metallurgy World
264 Philippe Toneguzzo et a!.
al.

Congress, PM' 94, vol. 3. Les Editions de Physique, Paris, p. 1823


(1994)
Koltypin, Yu.,
Yu. , G. Kataby, X. Cao, R. Prozorov and A. Gedanken. J. Non-
Cryst. Solids 201: 159 (1996)
Lalla, N. P. , M. S. Hedge, L. Dupont, K. Tekaia-Elhsissen, J. -M. Tarascon
and N. Srivastava. Current Science 78: 73 (2000)
Landolt-Bornstein. Magnetic Properties, New Series 111-190. 111-19a. Springer-
Verlag, Berlin, p.219
p. 219 (1982)
Legrand, J. , C. Petit, D. Bazin and M. -Po Pileni. Appl. Surf. Sci.164: 193
(2000)
Li, X. G., A. Chiba and S. Takahashi. J. Magn. Magn. Mater. 170: 339
((1997)
1997)
Li, X. G. , T. Murai, T. Saito and S. Takahashi. J. Magn. Magn. Mater.
190: 277 (1998)
Li, X.G. and S. Takahashi. J. Magn. Magn. Mater. 214: 195 (2000)
Li, Y.,
Y. , W. Gong, G.C.G. C. Hadjipanayis, C.M.
C. M. Sorensen, K.J.
K. J. Klabunde, VV.
Papaefthymiou, A. Kostikas and A. Simopoulos. J. Magn. Magn. Mater.
130: 261 (1994)
Lin, X. M., C. M. Sorensen, K. J. Klabunde and G. C. Hadjipanayis. J.
Mater Res. 14: 1542 (1999)
Mercier, D., J. C. S. Levy, G. Viau, F. Fievet-Vincent, F. Fievet, Ph.
Toneguzzo and O. Acher. Phys. Rev. B 62: 532 (2000)
Ocana, M., R. Rodriguez-Clemente
ROdriguez-Clemente and C. J. Serna. Adv. Mater 7: 212
(1995)
Oppengard, A. A.L.,
L. , F.J.
F. J. Darnell and H.C. Phys.Suppl.
H. C. Miller. J. Appl. Phys. Suppl 32:
1849 (1961)
Osuna, J. , D. de Caro, C. Amiens, B. Chaudret and E. Snoeck. J. Phys.
Chem.l00: 14571 (1996)
Ould Ely, T. , C. Amiens and B. Chaudret. Chem. Mater. 11: 526 (1999)
Overbeek, J. T. G. Adv. Colloid Interface Sci. 15: 251 (1982)
Petit, C. and M.-P. Pileni. J. Magn. Magn. Mater. 166: 82 (1997)
Petit, C., A. Taleb and M.-P. Pileni. Adv. Mater. 10: 259 (1998)
Puntes, V.F.,
V. F. , K.M.
K M. KrishnanandP. Alivisatos.
Krishnan and P. AI Appl. Phys. Lett. 78: 2189
ivisatos. Appl
(2001)
Respaud, M. , J. M. Broto, H. Rakoto, A. R. Fert, L. Thomas, B. Barbara,
M. Verelst, E. Snoeck, P. Lecante, A. Mosset, Mosse!. J. Osuna, T. Ould Ely,
C. Amiens and B. Chaudret. Phys. Rev. B 57: 2925 (1998)
Respaud, M. M.,, M. Goiran, J. M. Broto, F.
J.M. H. Wang, T. Ould Ely, C. Amiens
F.H.
and B. Chaudret. Phys. Rev. B 59: 3934 (1999)
Rousselle, D. , A. Berthault, O. Acher, J. P. Bouchaud and G. Zerah. J.
Appl. Phys. 74: 475 (1993)
Sapiezko, R. and E. Matijevic. Corrosion NACE 36: 533 (1980)
Shafi, K. V. P. M., A. Gedanken and R. Prozorov. Adv. Mater. 10: 590
((1998a)
1998a)
Monodisperse Ferromagnetic Metal Particles: Synthesis by...
by. . . 265

Shafi, K. V. P. M. , A. Gedanken and R. Prozorov. J. Mater. Chem. 8: 769


(1998b)
Shafi, K. V. P. M. , A. Gedanken, R. Prozorov, A. Revesz and J. Lendvai.
J. Mater. Res. 15: 332 (2000)
Shilov, V. P. , J. -C. Bacri, F. Gazeau, F. Gendron, R. Perzynski and Yu. L.
Raikher. J. Appl. Phys. 85: 6642 (1999a)
P. , Yu. L. Raikher, J.
Shilov, V. P., -C. Bacri, F. Gazeau and R. Perzynski.
J.-C.
Phys. Rev. B60:B 60: 11,902 (1999b)
Shinjo, T. , T. Shigematsu, N. Hosaito, T. Iwasaki and T. Takai. Jpn. J.
Appl. Phys. 21: L220 (1980)
Shipko, F.J. and D.L. Douglas. J. Phys. Chem. 60: 1519 (1956)
Sun, S. and C.B.
C. B. Murray. J. Appl. Phys. 85: 4325 (1999)
Tamura, I. and M. Hayashi. Surf. Sci. 146: 501 (1984)
Taylor, A. J. Inst. Metals 77: 585 (1950)
Toneguzzo, Ph., O. Acher, G. Viau, F. FiEwet-Vincent and F. Fievet. J.
Appl. Phys. 81: 5546 (1997) (1997>
Ph. , G. Viau, O. Acher, F. Fievet-Vincent and F. Fievet.
Toneguzzo, Ph., FiEwet. Adv.
Mater. 10: 1032 (1998)
Toneguzzo, Ph., G. Viau, O. Acher, F. Guillet, E. Bruneton, F. Fievet-
Vincent and F. Fievet. J. Mater. Sci. 35: 3767 (2000)
Uchikoshi, T., Y. Sakka and S. Ohno. In: Proc. 1st Japan International
SAMPE Symposium. p.25 p. 25 (1989)
Uehori, T.
T.,, A. Hosaka, Y. Tokuoka, T. Izumi and Y. Imaoka. IEEE Trans.
Magn. 14: 852 (1978)
G. , F. Ravel, O. Acher, F. Fievet-Vincent and F. Fievet. J. Appl.
Viau, G.,
Phys. 76: 6570 (1994)
Viau, G. , F. Ravel, O. Acher, F. Fievet-Vincent and F. Fievet. J. Magn.
Magn. Mater. 140-144: 6570 (1995)
Viau, G., F. Fievet-Vincent and F. Fievet. J. Mater. Chem. 6: 1047
(1996a)
Viau, G., F. Fievet-Vincent and F. Fievet. Solid State lonics 84: 259
(1996b)
Viau, G., F. Fievet-Vincent, F. Fievet, P. Toneguzzo, F. Ravel and O.
Acher. J. Appl. Phys. 81: 2749 (1997> (1997)
G. , Ph. Toneguzzo, A. Pierrard, O. Acher, F. Fievet-Vincent and F.
Viau, G.,
Fievet. Scripta Materialia 44: 2263 (2001)
Voltairas, P.A. andC.V. Massalas. J. Magn. Magn. Mater. 124: 20 (1993)
Voltairas, P. A. , D. I. Fotiadis and C. V. Massalas. J. Appl. Phys. 88: 374
0.1.
(2000a)
Voltairas, P. A. , D.
0.1.I. Fotiadis and C. V. Massalas. J. Magn. Magn. Mater.
217: Ll-4 (2000b)
Wei, D. and G.N. Patey. Phys. Rev. Lett. 68: 2043 (1992)
Weis, J. and D. Levesque. Phys. Rev. Lett. 71: 2729 (1993)
Wells, S. , S. W. Charles, S. M0rup, S. Linderoth, J. Van Wonterghem. J.
266 Philippe Toneguzzo et al.

Lasen and M. B. Madsen. J. Phys. Cond. Matter. 1: 8199 (1989)


Wernsdorfer, W. , E. B. Orozco, B. Barbara, K. Hasselbach, A. Benoit, D.
Mailly, B. Doudin, J. Meier, J. E. Wegrowe, J. -Po Ansermet, N.
Demoncy, H. Pascard, A. Loiseau, L. Francois, N. Duxin and M. -Po
(1997)
Pileni. J. Appl. Phys. 81: 5543 (1997>
Williamson, G.
G.K.
K. and
andW.H.
W. H. Hall. Acta Metal. 1: 22 (1953)
Yamada, H., M. Takano, M. Kiyama, J. Takada, T. Shinjo and K.
Watanabe. Adv. Ceram. 16: 169 (1985)
Yiping, L., G. C. Hadjipanayis, C. M. Sorensen and K. J. Klabunde. J.
Magn. Magn. Mater. 104-107: 1545 (1992)

Philippe Toneguzzo is indebted to Dr. Francois Guillet, CEA Le Ripault, for the XRD
investigations and Williamson and Hall analysis, and to Dr. Olivier Acher, CEA Le Ripault,
and to Prof. Olivier Tillement, Universite Claude Bernard Lyon I, for fruitful discussions.
8 Monocrystalline Half-Metallic NiMnSb Thin
Films: Preparation and Characterization

-Po Nozieres
Delia Ristoiu and J. -P.

8. 1 Introduction

The discovery of giant magnetoresistance (GMR) in magnetic multi layers


(Baibich et al.,
aI., 1988) and of tunnel magnetoresistance (TMR) CTMR) in metal/
insulator junctions (Julliere, 1975) has marked the beginning of spin
electronics, which makes use of the electron's spin as an additional degree of
freedom, thus enabling novel phenomena and devices to be imagined. The
GMR spin valves (Dieny et ai., al., 1991; Parkin, 1993) have already been
integrated in many functional devices Iike like magnetic read heads, angle and
position sensors and low-capacity random access memories (RAM). The TMR
tunnel junctions are mainly proposed for non-volatile magnetic memories
(MRAM) which combine all the advantages of conventional memories such as
DRAM (dynamic RAM), SRAM (static RAM) or flash: short access time, non-
volatility, insensibility to ionizing radiation, high density, infinite cyclabilty.
Major IC manufacturers are now ready to launch the first fully functional
prototypes.
Most of the devices developed so far are based on ferromagnetic 3d
metals such as Fe, Co, Ni and their binary and ternary alloys. The amplitude
of both GMR and TMR, however, is directly related to the detailed feature of
the band structure near the Fermi level. In particular spin polarization of the
conduction electrons is of primary importance in determining the amplitude and
the sign of magnetoresistance. For TMR junctions, the higher the spin
polarization of the two ferromagnetic electrodes, the larger the
magnetoresistance, hence the output voltage. Consequently, half-metallic
ferromagnetic materials, having fully spin polarized conduction band, are the
ideal candidates for such appl ications. Among the half-metall
applications. ic mater
half-metallic ials the
materials
most studied today are the Mn-based Heusler alloys (PtMnSb, NiMnSb), the
Mn-based perovskites (La (LaO? Sro. 3Mn03)
07 Sr03 Cr02 .
Mn03) and oxides such as Fe3 0 4 and Cr02'
Of the above, half-metallic semi-Heusler alloy NiMnSb appears the most
promising candidate thanks to its high magnetic ordering temperature (T c =
730 K) enabling room temperature functioning devices to be envisioned.
As band structure calculations have revealed that NiMnSb' shalf-metallic
268 Nozh~res
Delia Ristoiu and J. -P. Nozieres

character is strongly correlated to the structural and chemical order (Kulatov


and Mazin, 1990; Orgassa et al. , 1999), our attention is focused
focuse.d on epitaxial
thin films growth and on the related fundamental magnetic and transport
properties.
In the first section of this chapter, we will describe the main properties of
bulk half-metals and semi-Heusler alloys in order to throw in the basic
framework of the field. The second section is then focused on the growth of
epitaxial NiMnSb, while the third section is dedicated to the main transport
and magnetic properties of such films. In the fourth section we evidence a
crossover form a normal metallic-like behavior to a half-metallic-like behavior
at a crossover temperature around 100 K, and we finally will draw the
conclusions of this study.

8.2 The Half-Metallic Semi-Hensler NiMnSb

NiMnSb belongs both to the semi-Heusler alloys family for its crystalline
structure (Heusler, 1903), and to the half-metallic family for its electronic
structure (Groot et al.,
ai., 1983). In this section we will briefly review the main
characteristics of Heusler alloys and of half-metals in order to identify the
specificity of NiMnSb in the general framework of these two families. Finally
we will review the experiments dealing with evidence of the half-metal
character of this compound.

8. 2. 1 Heusler Alloys

In 1903 F. Heusler evidenced the ferromagnetic behavior of some ternary


"non-magnetic" metal-based alloys (the antiferromagnetism was unknown at
the time) such as CuMnAI and CuMnSn (Heusler, 1903). Later later studies showed
that this ferromagnetic order is associated with the onset of an ordered cubic
1951). These cub
phase (Bozorth, 1951) cubic
ic phases were called Heusler alloys.
The Heusler alloys, of chemical formulae X2 2 MnZ, where X is a transition

metal and Z is a sp type metal, crystallize in a face centered cubic structure


composed by four face centered cubic sub-lattices. The unit cell of an Heusler
alloy is represented in Fig. 8. 1. In this structure (L 21 ) belonging to the Fm 3m
(l21)
space group, the X atoms positions are 4a (0,0,0) and 4b (1/2, 1/2, 1/2), the
Mn atoms positions are 4d (3/4, 3/4, 3/4) and the Z atoms positions are 4c
( 1/4, 1/4, 1/4). An elementary cell is composed by 4 unit formula.
(1/4,
A new class of materials can be obtained by replacing the X atom on the
4a (0, 0, 0) site by a vacancy. These MgAgAs (C 1b ) type structures,
belonging to the F43m space group, are usually called semi-Heusler alloys.
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 269

/' .//' /'


/
/ ' ./'
/ /' X ,=Ni

-.
XJ=Ni
- -fA

~
.... /, 't-
-
If-
I

I
-
~
-"
:;~
, ,
,
I

, ./'
I
/'
o X 2=Vacancy
Mn
I

Y
.
I I
/ I V I ./" ~ Z=Sb
...
I
I ,
I //
fA-
-
JA-
-
f--l
f-----<
-4 '
.
./'
/ ./'
/' .
./'
/
/ / V
Figure 8. 1 Crystallographic structure of semi-Heusler alloys. Example for NiMnSb case.

From the very early studies on these materials the importance of the
crystallographic order on the physical properties was emphasized. R. M.
Bozorth (Bozorth, 1951) observed for example that the saturation
magnetization is strongly reduced if the Z and Mn atoms are randomly
distributed on the 4d and 4c sites.
Most of the Mn-based Heusler alloys are ferromagnetic, but some of
them, like CuMnSb are antiferromagnetic. These compounds are often cited as
examples of local ized magnetism (Hamzic et ai.,
localized aI., 1981; Moran et aI., 1994;
Rodriguez et al. , 1994) and their ground state is described by an Heisenberg
i#ci
i;ioj

type hamiltonian H
H = ~J ijS
=-- 2.:;J ijS is ij ,' where J ii
ij represents the exchange constant
i,j
i.j
between two Mn atoms situated on i and j positions; S i ' S ij represent spin of
atom i, j. However as the shortest Mn-Mn distances are of the order of
0.4 nm, the direct exchange interactions between Mn atoms are very weak
and the itinerant electrons of the X and Z atoms achieve the exchange
coupling. In order to better describe the magnetism of these materials, Kubler
introduced the notion of localized magnetism of delocalized electrons (Kubler,
1983) :"the
: "the localized
local ized moments are composed of itinerant electrons and result
from the exclusion of the minority spin electrons from the Mn sites".
Table 8. 1 Lattice parameter a (nm), valence electron number V. E. , Curie temperature
Tc(K) and magnetic moment per formula unit fJeff
Tc(K) (J.1B/ f. u.) for some semi-
J.1eff (fJB/

1997>
al. , 1997)
Heusler phases. (Pierre et ai.,
Phase V.E. a(nm) T c (K)
Tc(K) J.1eff(J.1B/f.
fJeff(fJB/f. u.
CoTiSn 17 0.5997 135 1.35
135
CoTiSb 18 0.5884
CoNbSn 18 0.5947
NiTiSn 18 0.5947
05947
CoVSb 19 0.5791
05791 11 -58
11-58 0.9-11.26
0.9- .26
NiTiSb 19 0.5872
NiMnSb 18 0.593
0593 730 45-2.9
PtMnSb 18 0.619 572 49-4.3
270 Delia Ristoiu and J.-P. Nozil~res
J. -P. Nozieres

Several studies concerning the semi-Heusler alloys XYZ (X=Ni, Co, Fe,
Y = Ti, Co, Mn, Z = Sb, Sn) have been undertaken recently (Kouakou et al. ,
1995; Pierre et ai.,
aI., 1997; Tobola et al., 1998; Kaczmarska et ai.,al., 1998,
1999) (Table 8. 1). It has been shown in particular that the number of valence
electrons per unit formulae governs the magnetic and transport properties of
al., 1997>.
these compounds (Pierre et ai., 1997). The structure holds through covalent
bonding, after transfer of s electrons from X and Y metals to the p shell of Z,
so the structure is most stable when the number of valence electrons is close to
18, favoring Sp3 hybridization around Z. Consequently, the semi-Heusler alloys
having 18 valence electrons like NiTiSn, NiZrSn, NiHfSn, CoTiSb, NiRSb
(R = Dy, Lu) are semiconductors, all the conduction electrons being involved
in the covalent bonding. Adding or removing one electron leads to a crossover
from a semiconducting to a metallic
metall ic system, with a broad spectrum of
magnetic order from Pauli paramagnet to Curie-Weiss behavior, and from
weak ferromagnet to strong half metallic ferromagnet. For example for 17
electrons we have CoTiSn (itinerant ferromagnetism), FeTiSb (Curie-Weiss
paramagnetism) and for 19 electrons we have NiTiSb (Paul(Paulii paramagnetism)
and CoVSb (itinerant ferromagnetism).
ferromagnetism) .

8.2.2
8. 2. 2 The Half-Metals

In 1983 R. A. de Groot (Groot et al.,ai., 1983) predicted by band calculations


that some Mn-based Heusler alloys should present a so-called half-metallic
band structure, characterized by a metallic
metall ic band structure for the majority spin
electrons (spin up) and a semiconducting band structure for the minority spin
electrons (spin down) (Fig. 8.2). In other words, this means that the Fermi
level crosses the sub-band of spin up electrons, and Iieslies in a gap of the spin
down electron sub-band. Therefore we deal here with a metallic material with
fully spin up polarized conduction electrons. Note that half metals should not be
confused with semi-metals like Sb, which exhibit a non-polarized, very small
density of states at the Fermi level.

A r !:l XZWDK L
WQ L Art:-. r A r !:l XZWDK L
XSU G L WQ L Art:-. r XSU G L
(a) (b)

Figure 8.2 Spin up (a) and spin down (b) NiMnSb band structure (Groot et al. , 1983).
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 271

Both the total number of electrons and the number of spin down electrons
being integers (because the spin down band is full), the spin up electron
number has to be an integer too. As a consequence, the magnetic moment of a
half-metal, if we suppose that there is no polarization of the non-magnetic
levels, is also an integer.
The half-metallic character was predicted in several Mn-based Heusler
(4!Js If. u. ,where !Js
alloys amongst which NiMnSb and PtMnSb (4J..1s/f. J..Is is Bohr
magnetron, f. u. is formula unit) are the most widely known (Groot, 1991).
CoMnSb should also be a ferromagnetic half-metal with a moment of 3 !Js/f. J..Is/f. u. ,
FeMnSb a ferromagnetic half-metal with a moment of 2 !Js/f. J..Is/f. u. , MnMnSb a
J..Is If. u. and finally CrMnSb an
ferromagnetic half-metal with a moment of 1 !Js
antiferromagnetic half-metal with Mn and Cr moments canceling
cancel ing each other.
Other semi-Heusler alloys like TMnSb (T = Rh, Pd, Ir, PO could be half-
metallic under high pressure (Ishida et ai.,al. , 1997).
Xz2MnZ type full Heusler alloys were also investigated by band structure
calculation by Ishida et al. (Fujii
(Fuji i et al. , 1990; Ishida et al. , 1995a, 1995b) in
1995a,1995b)
order to evidence a half-metal character. Compounds like CO COz2 MnZ (Z = Si,
Ge), Fez
Fe2 MnSi and Ruz
RU2 MnZ (Z = Si, Sb) were found to have a half-metallic
band structure in the ferromagnetic state.
The half-metallic character is not only related to Heusler or semi-Heusler-
based structures (Table 8. 2). It was also predicted in Cr02CrOz (Schwarz, 1986),
Fe304
Fe3 0 4 (Yanase and Siratori, 1984) and some Mn-based perovskites like
LaQ7XQ3Mn03(X=Ca,
La 07 X03 Mn03(X=Ca, Sr, Pb) (Pickett and Singh, 1996).

Table 8.2
8. 2 The Curie temperature T cc CK),
(K), saturation magnetization /.10 M s(T),
CT), the effective
magnetic moment per formula unit /.Ieff (/.Ie/
C/.Is/ f. u.) and the crystallographic
structure of some half-metallic materials.

Compound T c (K)
CK) /.IoMs(T)
/.10 MsC T) /.Ieff (/.Ie/f. u.
/.IeffC/.Is/f.u. Crystallograpic structure

La07Sr03 Mn03
Lao.7 8ro. 3Mn03 350 0.74 3.7 Perovskite

NiMn8b
NiMnSb 728 0.89 4 8emi-Heusler
Semi-Heusler

Cr02 396 0.8IT


0.81T 2 Rutile

Fe 3O.
Fe30, 860 0.63T 4 Inverse spinel

PtMn8b
PtMnSb 572 0.9 4 Semi-Heusler
8emi-Heusler

Sr2 FeMo06
8r2FeMo06 415 0.73 4 Double perovskite

Where /Jo is vacuum magnetic permeability, M s is saturation magnetization.


magnetization, /Jeff
/Jeff is the effective

magnetic moment per formula unit.


unit, and Tc
T c is Curie temperature.

Note that all these predictions are based on band structure calculations of
the ordered compound at zero Kelvin. Direct evidence of the half metallic
272 Delia Ristoiu and J. -P. Nozil!res
Nozieres

character is troublesome, and most experiments lead to controversial results.

8.2.3
8. 2. 3 The Case of NiMnSb

The intermetallic compound NiMnSb was synthesized for the first time by
Castelliz in 1951 CCastelliz,
(Castelliz, 1951). NiMnSb crystallize in a C 1b type (semi-
Csemi-
O. 5901 nm.
Heusler) face centered cubic structure with a cell parameter of 0.5901
This compound is ferromagnetic with a Curie temperature ranging between
730 - 750 K, according to different authors (Endo, 1970; Webster and
Mankikar, 1984; Helmholdt et al., 1984; Otto et al., 1987), and a low
temperature saturation magnetic moment very close to 4 I-is
IJs per formulae unit.
The neutron diffraction studies (Webster and Mankikar, 1984; Hordequin
et al. , 1997a) have shown that this magnetic moment is mainly confined on
the Mn atoms (3. 8 I-is)'
IJs)' with only a low contribution associated to the Ni
atoms (0.2 I-is)
IJs) and no contribution coming from the Sb atoms.
The NiMnSb electronic band structure calculated by de Groot in 1983
(Groot
CGroot et al. , 1983) reveals a 0.5 eV gap in the spin down density of states,
with the Fermi level lying approximately in the middle of it. Relativistic band
structure calculations for NiMnSb and PtMnSb by Youn et al. (Youn and Min,
1995) confirm the half-metall ic character only for NiMnSb. Kulatov et al.
half-metallic
CKulatov and Mazin, 1990), however, point to the limitations of band structure
(Kulatov
calculations, in particular, concerning the gap value which seems to depend
strongly on the atomic size and the exchange term chosen.
The consequences of the atomic disorder on the different crystallographic
sites were recently theoretically treated by Orgassa et al. (Orgassa
COrgassa et al. ,
1999). The existence of a vacancy in the half-Heusler structure favors some
local atomic disorder whose consequence must be taken into account. The
calculations undertaken by Orgassa et al. show that an atomic disorder as
small as 5 % is enough to destroy the half-metallic character. The authors
identify 3 different types of disorder: CD interchange of Ni and Mn atoms,
CV equal amounts of Ni and Mn on the vacancy site and @ equal amounts of Mn
and Sb on the vacancy site (Fig. 8.3).
8 3). For the first 2 cases a disorder level of
5 % still preserves the spin down gap, but the gap does not contain the Fermi
closed.
level anymore, while in the 3rd case the gap is completely closed
Many experimental studies were undertaken in order to prove the half-
metallic character of NiMnSb. More often spin resolved spectroscopy
techniques were undertaken to directly probe the spin up and spin down
density of states and evidence the gap. First Bona et al. (Bona et al. , 1985)
studied a polycrystalline NiMnSb sample by spin resolved photoemission, but
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 273

-------,------------,

1.0
.:::::---
.:::::-- _
-- -.:-::-
~ - :::-=--=
A-B

=-:.:..-=-----
- --..:-.: =-=.:..
~
c::
.~
--
0.5
">
<.;)

~
</>

~ ,,/ /
/::,
~
~ :,,
~
-:---.._-- --1%'
,

-1.0 - - -6% :
_- ..- -10%:
- 100/0;
,
..,-E-,Il_--,-L_-:-'-:-_c:-"L--:'--:-'
L...j'"-:-_-J-,--:'-'
-OJ
-0.3 -0.2 -0.1 0o 0.1 0.2 0.3 0.4
E(eY)
(a)

c
c:
c::
.~
'~
0.5
">
0;
u
<.;)
Mn on C

~ 0 *~~
,
---+-,---.--'-'-'-~-=r---+i":--:'-=;
~
~
Vl
t/l
:=:: -0.5
.::::
- -..........
--.......................
...................... :
,
~
'::- -: .............. >--...:
~: ,
"

~ -1.0 ~
,,~--
,
---- ---- --10/0
- - -5%
- 50/0
----100/0
- -10%
I....
I...

:'
:
, ,
-0.3 -0.2 -0.1
-0.\ 0o 0.1 0.2 OJ
0.3 0.4
E(eY)
(b)

1.0 -- ---~
C'

c?
.~ 0.5

">
0;
u
<.;)

~
~
~
~
;'l
~ -0.5
-'' ....
, ..... - ----...-... ....... ,:
,
:
Si
~ -1.0

-0.4
-7--- ___
-~---
:--1% -
: - - -5%

-0.3 -0.2 -0.1


- 5,10
: - - --10%
100/0
0
EF
--- ',,:--'"':
----
0.1 0.2 0.3
,
::
,'

E(eY)
(e)

Figure 8. 3 Calculated spin-polarized electronic density of states N H (E) for


NiMnSb with Ni-Mn site disorder (a), NiMn-vacancy site disorder (b) and MnSb-
vacancy site disorder (c) respectively. Disorder levels are 1 % (solid lines) 5 %
(dashed lines) and 10% (dash-dotted lines) (Qrgassa
COrgassa et ai., 1999).
274 NoziEHes
Delia Ristoiu and J.P. Nozieres

they found only 50 % spin polarization at the Fermi level at low temperature
(20 K). They concluded that if NiMnSb was to be really half-metallic, the spin
down band gap should be less than 0.5 eV, which represents the experimental
resolution of the spectrometer used. This result must however be interpreted
with care as: CD the photoemission is a surface analysis and the NiMnSb
surface is complex (ternary compound with a vacancy site) and consequently
prone to off-stoechiometry or segregation; (2) the sample used is
polycrystall ine.
The first experimental proof of the half-metallic character of NiMnSb at
low temperature (10K) was obtained by spin resolved positron annihilation
measurements undertaken by Hanssen et al. (Hanssen et al., 1990). The
volume probed by this technique is several microns, so the result obtained
characterizes the bulk. Note, however, that the theoretical approach required
for interpreting the experimental data is rather complex (Hanssen and
Mijnarends, 1986) and thus the conclusion is quite indirect
indirect.
Other indirect evidences of NiMnSb 's half-metallic character were
obtained by resonant photoemission (Robey et al., 1992), 1992) , infrared
reflectivity (Mancoff et ai.,
al., 1999) or inelastic neutron scattering (Hordequin
et al. , 1997b). For all these studies, the half-metallic character is probed by
comparing the experimental spectra to the theoretical predictions.
A different approach, although still indirect, consists in measuring the spin
polarization of the conduction electrons by Andreev reflection at a metal/
superconductor interface. In the ideal case of a half-metal/superconductor
interface, the current should not pass through the interface, for the Cooper
pair formation is impossible. Soulen et al. measured the spin polarization of
many ferromagnetic materials (Ni, Co, Fe, NiFe, NiMnSb, Lao? Lao,? Sr03 Mn03,
Mn03 ,
CrOz)
Cr02) using a superconducting tip (Soulen et ai., al., 1998, 1999). The NiMnSb
spin polarization measured this way was only 58 % at 1. 6 K. Nevertheless,
the results of these measurements may again be affected by the surface
quality,
qual ity, which was neither specially prepared nor characterized.
Finally, Ristoiu et al. (2000a) undertook a detailed surface study on
epitaxial NiMnSb films. For a monocrystall ine stoechimetric surface, the spin
polarization at the Fermi level, directly measured by inverse spin polarized
photoemission, is found to be 100 % above the background at r point (k = 0) =
and at room temperature. More recently, similar measurements performed
over the full Brillouin zone revealed a polarization of less than 75 % (Turban,
2001).
2001 ). The stoechiometric surface, however, was found to be extremely
fragile, as classical surface treatments like sputtering or high temperature
anneal led to Mn segregation and complete spin depolarization of the
conduction band (Ristoiu et al. , 2000b). This extreme surface sensitivity may
explain the low values of spin polarization obtained when starting with
polycrystalline films or uncharacterized surfaces.
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 275

8.3 Growth and Structure of Epitaxial NiMnSb Thin Films

8.3.1
8.3. 1 PtMnSb

The studies of half-metallic Heusler alloy thin film deposition started with
PtMnSb because of its large Kerr rotation (Engen et al., 1983; Groot and
Bushow, 1986) which can be used for magneto-optical recording. Three
fabrication methods to obtain polycrystalline single phase PtMnSb thin films
can be distinguished: CD low temperature deposition 100 c ) followed by a
crystallization vacuum anneal at high temperature (typically 500C)
500 C) (Ohyama
et al., 1985, 1987; Attaran and Grundy 1989), CZ) ~ direct crystallization by
high temperature deposition at 200-500
200 - 500 c (Shoji et a!.,
al. , 1987; Naoe et a!.,
al. ,
1988, 1994) and @ sequential deposition of the three elements in equimolar
quantities followed by a high temperature anneal (Inukai et al., 1986).
Sputtering (rf diode, Ohyama et al. a!. ,1985, 1987),
1987>, rf magnetron (Takanashi
<Takanashi
et al., 1987; Shoj
Shojii et al.,
aI., 1987), facing targets (Naoe et al.,aI., 1988) and
evaporation techniques (Inukai et al., 1986; Shiomi et al., 1988) were
successfully used and the films obtained on glass or Si02 were polycrystalline.
Later the group of B. M. Clemens reported for the first time on epitaxial
PtMnSb obtained by dc de magnetron co-sputtering of the constituting elements at
500C,
500 c , using monocrystall ine MgO (001)
monocrystalline (00 1) substrates covered by a W buffer
layer (Kautzky and Clemens, 1995; Kautzky et aI., a!. , 1997>.
1997).

8. 3. 2
8.3. PtMnSb/NiMnSb Multilayers
Takanashi et a!.
al. (1987>
(1987) prepared PtMnSb (5 nm) /NiMnSb (5 nm) multi layers
using a dual-type rf magnetron sputtering method. The multi layers are deposed
on a glass substrate at 50C
50 c and under 10 mTorr Ar pressure, followed by a
500 c post deposition anneal. (111) textured super-lattices were obtained
500C
and the Kerr rotation measured was intermediate between that of PtMnSb and
that of NiMnSb.
Bobo et a!.
al. (1997 a) fabricated (111) epitaxial superlattices of XMnSb
(1997a)
(X=Cu,Ni) /PtMnSb on (0001) AI 2 0 3 substrates. A first epitaxial PtMnSb
Ah03
layer deposited at 500C
500 c is used as a buffer for the subsequent superlattice:
superlatticE:
growth at lower temperature (200 - 500C)
500 C) (Fig. 8. 4). For multilayers
deposited at high temperature (350 - 500C)
500 C) additional phases are obtained
(MnSb), but for an optimum deposition temperature of 300C
300 c a good epitaxy
is evidenced. Note that PtMnSb/NiMnSb superlattices present a tendency
276 Nozil~res
Delia Ristoiu and J.-P. Nozieres

towards a perpendicular anisotropy, due either to the lattice mismatch or to


the interface anisotropy.

Buffer Superlattice
~ 104 (III)
(I I I) (111)
(III)
"
<l)

'"~ 10
u
()

0
3

v
.~
i:'
'Vi 102
c:
c::
2~
c:
c::
10 11

16 18 20 22 24
28(")
2en
(b)

Figure 8.4 (a) Grazing incidence X-ray (Cu-K a1 wavelength) diffraction around the
[200J direction for the PtMnSb buffer and the NiMnSb (9.4(9 4 A) (9. 4 A)
A)I/ PtMnSb (9.4
superlattice;
superlatlice; (b) Symmetric diffraction at 10 keV on the same superlatlice
super lattice around the
111 peak compared to the kinematics simulation (Sobo
<Bobo et al. , 1997a)
1997a)..

8.3.3 Polycrystalline NiMnSb Thin Films

The advantages of using fully spin-polarized materials for spin electronics


applications enhanced the interest in NiMnSb thin film fabrication for
implementation in magnetic heterostructures. The main reasons for using
NiMnSb instead of PtMnSb is the higher Curie temperature 030
(730 K with respect
to 580 K for PtMnSb), the lower fabrication costs and the low spin orbit
coupling (which is neglected in the band structure calculations), which may
affect the half metallic character.
aI., 1990) reported for the first
The group of J. S. Moodera (Kabani et al.,
time on NiMnSb films fabrication by co-evaporation of the individual elements
Substrates were either glass or [0001] sapphire heated at 350 'c "C during film
deposition. These films are then annealed in situ at 500 'c
"C for 2 h and then
nitrogen quenched. Finally the films obtained are polycrystalline on glass and
(111) textured on sapphire. The lattice parameter (0.5907 nm) is very close
to the bulk value and the film composition is stoichiometric. The magnetic
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 277

measurements reveal an in-plane anisotropy and a magnetic moment of (3.9


O. 2) J.1s
JJB per formula unit, in agreement with the bulk. Transport studies
undertaken on these thin films (resistivity, magnetoresistance and Hall effect)
(Moodera and Mootoo, 1994) show an anomalous thermal behavior around 100 K
(Fig. 8.5), as already observed on bulk NiMnSb by Otto et al. (1989a,
1989b ) . The decrease in resistivity, the positive (cyclotron)
magnetoresistance and the Hall coefficient increase below 100 K suggest an
increase of the electron mobility below this temperature, which could possibly
be due to a half-metallic low temperature state.

o
o
o
70 o 0
o
E NiMnSb 0 0
00
9s::,
I
0
0
:::;:: 00
00
0
0
0
2S 0 0
c
0
.;; 0
0
0
:~
'~ 0
.;;:;
'v;
0)
<I)
0
e:::
Cl':
50 0 PtMnSb
0
0
0
0

o 100 200 300


Temperature (K)

Figure 8.
8.55 Temperature variation of resistivity for NiMnSb and PtMnSb thin films.
Note the rapid drop in resistivity below -100
- 100 K for NiMnSb (Moodera
CMoodera and Mootoo,
1994) .

The determination of the NiMnSb spin polarization by measuring the


superconductor (AI) to ferromagnetic (NiMnSb) tunnel current through an MgO
barrier was unsuccessful, as no asymmetry of the conductance versus bias was
observed. The Auger study of the NiMnSb/MgO interface evidenced the
formation of MnO x ' yvhich may be responsible for the total depolarization of
the tunnel current.
Another approach to measure the spin polarization is to study the
magnetoresistance of a ferromagnetic/ insulator/ferromagnetic tunnel junction.
ferromagnetic/insulator/ferromagnetic
Good NiMnSb/
NiMnSb/AI AI interface quality was achieved by in situ low temperature
(77 K) AI deposition followed by AI oxidation by glow discharge in O2
278 Delia Ristoiu and J.-P.
J. -P. Nozieres
NoziEHes

atmosphere (Tanaka et aI., al., 1997>. Tunnel junction using CoFe and NiFe as
the second electrode were fabricated, but the obtained tunnel
magnetoresistance (TMR) was low (3.7% at 300 K and 8.1 % at 77 K) with a
corresponding calculated NiMnSb spin polarization of only 15 %. These results
were explicated by considering possible interface damage by oxidation and/or
segregation, and the necessity of in situ surface studies (XPS, Auger) was
put forward.
Later, two other groups, J. A. Caballero et al. (1997) (Childress et al.
J.A. aI.,,
1997) and C. Hordequin et al. (1998), obtained good quality polycrystalline
films on glass and monocrystalline (001) Si substrates by direct crystallization
at relatively low deposition temperatures (250 - 300 'C). J. A. Caballero
et al. (1997> (Childress et al., aI., 1997> prepared the films by rf magnetron
sputtering of a single composite NiMnSb target in an Ar atmosphere of 1.5 - 5
mTorr. Low sputtering power and low deposition rates favor NiMnSb
crystallization at relatively low deposition temperatures, but a parasite NiSb x
phase is evidenced. Lowering the Ar pressure during the deposition is required
to obtain single-phase stoichiometric NiMnSb films. The lattice parameter
(0. 5907 nm) and the magnetic moment per formula unit (4. 2J.Js 2/Js O. 2/J
2J.J s) are
in agreement with the bulk values. The anomaly in the temperature
dependence of the resistivity around 100 K is also observed.
The authors used these films in NiMnSb/Cu/NiMnSb/FeMn spin valves
(Caballero et al.,
al. , 1998a, 1999), but the obtained magnetoresistance both in
CIP and CPP geometry remained low (-- (,,- 7 % at 4. 2 K). As it was pointed
out, a poor interface quality might explain such low GMR amplitude.
Simultaneously, C. Hordequin et al. (1998) prepared NiMnSb films by
facing targets sputtering (FTS) using a pair of stoichiometric NiMnSb targets.
The best crystalline quality was obtained by depositing onto monocrystalline Si
substrates at 270 - 300 'c and under 10 mTorr Ar pressure. Similarly, low
deposition rate and low argon pressure favor a better crystalline quality. The
films are polycrystalline and stoichiometric and the lattice parameter (0.592 (0. 592
nm) as well as the magnetic moment (3. 8J.Js 8/Js O. 4/Js)
4J.Js) are in agreement with
the bulk. Note that the residual resistivity, 20 IJQcm,
I-lacm, is much lower than that
previously reported both by Moodera et al. (52 I-lacmIJQcm (Moodera and Mootoo,
1994
1994 and by Caballero et al. (1998b) (54 IJQcm). I-lacm). This low residual
resistivity is a signature of an improved crystalline quality. The transport
properties were investigated and the same anomaly in the thermal behavior of
the resistivity was evidenced around 80 K.
NiMnSb/(Mo, Cu)/NiMnSb/SmC0 2 spin valves were then fabricated in
order to evidence the half-metallic character of NiMnSb, but the GMR values
again remained below expectations ('" ("- 1% ), most probably due to
interdiffusion at the (Mo, Cu) /NiMnSb interface. This was further confirmed by
(Mo,Cu)/NiMnSb
the large magnetic dead layer of NiMnSb and the large resistivity at low
thickness of Mo.
All these studies on NiMnSb polycrystalline films point out the importance
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 279

of both the interfaces and of the crystalline quality on the spin dependent
transport properties, hence the interest in epitaxial NiMnSb films has
increased recently.

8.3.4 Epitaxial Growth of NiMnSb

J. F. Bobo et al. (1997b) fabricated the first epitaxial NiMnSb films by dc


magnetron co-sputtering. Using (0001) AI Al z2 0 3 substrates and deposition
temperature around 500'C, they obtained single phased (111) epitaxial and
stoichiometric films. Transmission, reflectivity and ell ipsometry spectra were
taken in order to determine the band gap in the minority spin band. These
results are consistent with the existence of a band gap of 1. 1 eV for NiMnSb.
No further studies were carried out.
J. S. Moodera et al. (Tanaka et aI., ai., 1999) fabricated (001) epitaxial
NiMnSb films by MBE onto (001) MgO substrates using a V buffer layer and a
deposition temperature of 473 K. The films obtained were used in tunnel
junction structures (NiMnSb/ Aiz Ab 0 3 / AI and NiMnSb/ Aiz Ab 0 3 /NiFe ) . The
measured TMR was higher than for polycrystalline junctions (19.5 % at 4.2 K
here), but still below expectations for a half-metallic system. The value for
the NiMnSb spin polarization deduced by superconductor to NiMnSb tunneling
was 28
28% % and by ferromagnet to NiMnSb tunneling
tunnel ing was 25%.
25 %. Possible NiMnSb
interface oxidation may be responsible for these law values of the spin
polarization.
van Roy et al. (2000) fabricated (001)
(DOl) epitaxial NiMnSb films onto GaAs
(001) substrates by MBE in order to use this system for spin injection devices.
In order to prevent GaAs/NiMnSb interface mixing, interface layers as AlAs,
GaSb and AISb were used. Single-phase epitaxial and stoichiometric films
were obtained for an optimum deposition temperature of 300 'c .
Ristoiu et al. (2000c) have fabricated stoichiometric NiMnSb thin films by
facing targets sputtering at high substrate temperature (400 - 500 'C) using
1 : 1 : 1 Ni : Mn : Sb targets. The substrate used is again monocrystalline
(001) MgO. A Mo buffer layer was required to achieve a good epitaxy. The
optimum deposition conditions (temperature, Argon pressure) were
determined to achieve the structural and chemical properties required for spin
electronics devices fabrication.
More recently, epitaxial (111) NiMnSb films were also fabricated by
molecular beam epitaxy (MBE) under UHV (Turban, 2001). Again an (001)
MgO substrate was used as well as a buffer layer of V. The growth was
performed by co-focal deposition of pure Ni, Mn and Sb elements. The
deposition conditions were optimized to achieve truly 2-dimensional growth.
The UHV conditions allowed for a post deposition high temperature anneal to
minimize surface roughness.
In this review, we discuss mostly the results obtained on the two latter
280 Delia Ristoiu and J.-P. Nozithes
J. -P. Nozieres

systems. In the following, unless otherwise stated all data deal ing with Mo
dealing
buffer refer to our FTS sputtered films, whereas data dealing with V buffer are
MBE-grown films of (Turban, 2001).
1. Epitaxy onto MgO (001)
MgO is a face-centered cubic crystal as NiMnSb, with a cell parameter QMgO =
aMgO =
0.4213 nm. Supposing that NiMnSb grows such that the [110 J J MgO axis is
parallel to the [100J NiMnSb axis (a MQo.J2 ~ a NiMnSb ), the lattice mismatch is
CQMgOJ2"""'QNiMnSb),
only 1 %, which is very promising for epitaxy.
It is generally difficult to predict the growth mode of a metal onto a
substrate by considering only surface energies. This is due to CD the surface
energy values for thin films being possibly different from the bulk, due to
modifications of the free surface electronic structure and (2) the surface
energies of complex compounds being badly known even for bulk crystals. The
direct epitaxy of NiMnSb onto MgO leads to the formation of isolated islands
and has been found to be impossible, whatever the deposition temperature.
Schlomka et al. (2000) also reported on the non-epiatxy of NiMnSb on MgO.
2. Epitaxy onto MgO/Mo or MgO!V
From the above, a buffer layer with good wetting properties has to be
intercalated between MgO and NiMnSb. Body-centered cubic Mo, with a cell
parameter a = 0.3147 nm ("""'QNiMnSb/
Q Mo =0.3147 (~aNiMnsb/2),
2 ), is one of the metals presenting the
higher surface energy (Bauer and Merwe, 1986). The 8 - 28 X-ray diffraction
diagram obtained on a film grown at 300C
300C with an Ar pressure of 10 - 2 Torr
on a Mo buffer layer, is shown in Fig. 8.6. 8. 6. For both Mo and NiMnSb, only
hOO peaks, with h even (h odd is forbidden by the extinction rules) are
observed. The calculated lattice parameter is (0. 59 o. O. 01) nm, very close
to the value reported for bulk single crystal (Castelliz, 1951). The crystalline
order in the planes perpendicular to the film surface can be investigated by
grazing incidence X-ray diffraction on a 4 circles diffractometer.
diffractometer The scans of

(j
i::S
::.::
:><:
.D
Vl
6on
0Of)
c:
<::
2
:2 2
:2
0
i 0
N
~

::i
~
0
0
N
=
::.::
:><:
0
2
:2
0
~ 6on
0 0
S
~
on
Of) 2
00
:2
N

.3
..3 0
0
N

20 30 40 50 60
2ee))
28("

Figure 8.6
8. 6 X-ray diffraction diagram (i\CuKa o. '542 nm) on a NiMnSb 50 nm
o.CUKo = 0.1542 I 10
'0 nm
Mol MgO film. "* "corresponds to the sample-holder peaks.
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 281

the 200 MgO, Mo and NiMnSb peaks in an angular range of 120 are reported
in Fig 8. 7. The 200 diffraction peaks of both Mo and NiMnSb have a 90
periodicity, as expected for a 4-fold symmetry. The mosaicity calculated from
the peak width is 1.5 for Mo and 2.5 for NiMnSb.
14

12


~ 10
--:-
I'M"Sb I

~.,.,
~.,.~_, ..".L
:::l

l
,;
oj
'0 8
x
2S

III ht1l'N
,p,: . _"'I
l'
0
.;;;
6
I
c
I Mo
2:l
c 4
II


I
I
a
0
2
,J
I JMgO
!MgO
I !
J
-'-----
..L-- I
-100 -50 a0 50 100 150 200 250
ep-xC)
<p-xC)

Figure 8. 7 Grazing incidence X-ray diffraction diagram for the 200 NiMnSb peak,
200 Mo peak and 200 MgO peak.

The 200 Mo peaks are shifted by 45 with respect to the 200 MgO peaks,
while the 200 NiMnSb peaks are aligned with those of MgO. This means that
the epitaxy is such that Mo has its [100 J axis rotated by 45 with respect to
the [100J axis of MgO, while the growth of NiMnSb onto Mo is cube on cube
([100J NiMnSb axis parallel to [100J Mo axis).
These epitaxy relations can be simply explained by considering the
relations between the lattice parameters of these 3 materials. In a (001) plane
a Mo cell can grow on the MgO cell diagonal (J2 OMgO::::::::: OMO)' with a lattice
OMgO ""'=' OMO)'

mismatch of 5.2 %. In the same plane, a NiMnSb cell can grow on top of 4 Mo

i;{
cells (2o Mo ""'='ONiMnSb), with a lattice mismatch of5.8%
(20Mo:::::::::ONiMnSb)' of 5.8% (Fig. 8.8).

~
aN"iMnSb ~

Q
GN,tMnSb
~"'"" 0 Mo
o 0,0 (Ni,Vacancy)
0,0 (Ni,Vacancy) /1:Mo
. / {.
or(Mn,Sb)
or(Mn,sb) /Ci
/a Mo

Figure 8.8
8. 8 Schema of the epitaxy relations between NiMnSb, Mo and MgO, in a
COOl) plane.
(00l)

Similarly, V buffer layers can be used and lead to epitaxy as good as with
Mo layers. V is a body centered cubic crystal with a lattice parameter of Ov =
Ov =
O. 303 nm. Hence the lattice mismatch with NiMnSb is 2.5 %. Extensive Auger
0.303
282 Nozil~res
Delia Ristoiu and J.-P. Nozieres

analysis revealed that the interdiffusion of V into NiMnSb is negligible when the
temperature is kept below 600'C600C (Turban, 2001). Typical RHEED pattern at
the different stages of growth are reported in Fig. 8.9. The epitaxy relations
[1 OOJNiMnsb II [100Jvll
are found to be [100JNiMnsbll [1 OOJv II [110JMgo.
[110 JM90 .

Figure 8. 9 RHEED patterns obtained on MgO (00')


(001) surface, V (001) surface and
NiMnSb (001) surface respectively (Turban,
<Turban, 2001).

8.3.5 Surface Morphology of the Epitaxial Films


1. Morphology of the buffer layer
To integrate these films in devices (spin valves, tunnel junctions) the
crystalline quality is not enough. The surface planarity is equally essential in
order to avoid pinholes in the non-magnetic barrier or magnetostatic "orange
peal" coupling
coupl ing between the two magnetic layers through seed barrier.
Moreover, a low quality surface or interface can lead to the decrease of the
magnetoresistive properties through spin-independent scattering (Moodera and
Mathon, 1999) and eventually to the depolarization of the conduction electrons
by spin flip
fl ip scattering or by depolarization of the surface itself.
The 10 nm Mo buffer layer deposed at 450 450C'c on MgO presents a
continuous surface of average roughness (0. 5 0.2) nm (Fig. 8. 10) .
2. Morphology of the NiMnSb layers
For a 50 nm NiMnSb layer grown onto a Mo buffer, at 450 450C'c and under
10- 2 mbar (1 bar= 10 5 Pa) argon pressure, we observe that the film is highly
discontinuous with furrows as deep as the film thickness (Fig. 8. 11 lla).
a). The
furrows are oriented parallel or perpendicular to the [110][110J direction, cutting
up the film in almost square islands. The island surface is quite plane, with a
roughness of 0.5 - O. 7 nm rms, equivalent to that of Mo. Such a behavior is
specific to a 3-dimensional growth with random nucleation on the surface
defects of the buffer layer. A tendency towards coalescence is observed when
the film thickness increases as can be seen in Fig. 8. 11 b. Such coalescence is
characterized by an exponential growth of the island surface, but the
preferentially oriented furrows still exist.
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 283

A
600

400

200

o 4 8
(~lIl1)
(pm)

Figure 8.10 ~m x 10 ~m
8. 10 Image of a 10 IJm IJm region of a 10 nm Mo/MgO surface.
The two points of high contrast represent the tip unhooks on two dust particles.

A
1200

gOO
800

400

or

"~.~"
t,flll
.I\I(}
':111)

""c "
'"
~
~) 0.1
E
<" /
0.01
001
0'-------2-'-0---4--'O---6'-0---g-'-0---I--'O-0--1-J
0'---2-'-0,------4--'0---6'-0---8-"-0---1--'0-0--1-'20
20
thickl1~ss (11m)
Film thil:kIlCSS(IlIll)
(bl
(b)

Figure 8. 11 (a) AFM image of a 10 ~m ~m NiMnSb surface deposed at


IJm x 10 IJm
450 'c under 10-
450'C 10 - 22 mbar Ar; (b) The average island surface (S) as a function of
film thickness (t). The films thicknesses are 10, 50 and 100 nm respectively.
284 Delia Ristoiu and J.-P. Nozil~res
J.P. Nozieres

In order to obtain a 2-dimensional growth, the deposition parameters can


be varied, in particular the substrate temperature and the argon pressure. As
the deposition temperature at constant argon pressure is increased (Fig. 8. 12a)

if'!
~r".' If
:~ ~ I>

12~~
800
';iI
...
1
400
~1lIl ~d
o l~

Y ...
I I
.~ ~~ I
o 4 8
(J.1I11)
(a)

6to~
400

200
o

6:a~
6to~
400

200
o

I I
4 8
(J.1I11)
(!lm)
(e)

Figure 8. 12 NiMnSb surface morphology as a function of the deposition conditions:


(a) 10 IJm x 10 IJm NiMnSb surface deposed at 550 C 'C and 10- 2 mbar Ar; (b) 10 IJm x
10 IJm NiMnSb surface deposed at 450 'cC and 3 x 10- 3 mbar Ar; (c) 10 IJm x 10 IJm
NiMnSb surface deposed at 400 C 10 - 33 mbar Ar.
'C and 3 x 10-
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 285

better defined square islands are observed. This feature can be simply
understood by taking into account nucleation and growth thermodynamics in
sputtered films (Lewis and Anderson,
Anderson. 1978): higher incident energies lead to
a larger mobility of the ad-atoms which can diffuse to the existing nucleation
centers, therefore promoting a 3-dimensional growth.
Although not as dramatic as the deposition temperature.
temperature, the argon
pressure also plays an important role for the sputtered films morphology.
Decreasing the argon pressure (Fig. 8. 12b) leads to more continuous films,
films.
with islands close to coalescence. It is generally expected in the case of
sputtering growth (Thornton,
(Thornton. 1982) that higher argon pressure promotes a 3-
mobility,
dimensional growth by reducing the kinetic energy, hence the mobil ity, of the
ad-atoms.
Consequently both the deposition temperature and the argon pressure
have to be decreased in order to achieve a 2-dimensional growth (Ristoiu
et al.,
al.. 2000c). The best conditions are thus given on one hand by the
limitations related to the deposition chamber (argon pressure higher than 3 x
Iimitations
10- 3 mbar in order to obtain a stable plasma) and on the other hand by the
growth of NiMnSb (temperature above 400 'c to achieve a good epitaxy). The
surface of such a film (Fig. 8. 12c) is clearly smoother,
smoother. with an average
roughness of (1 0 . 2) nm rms.
The surface morphology of MBE-grown films onto a V buffer layer after a
675'C planarization anneal shows terraces of several hundreds of nm with
edges preferentially along [110]
[110J and [1-10]
[1-10J directions. The depth of these
steps is around 1 - 2 nm. On each terrace, several surface reconstruction
patterns are observed (4 x 4, 2 x 2) depending on the planarization annealing
(Turban, 2001). As can be seen on the TEM image of MBE-
temperature (Turban.
grown V/NiMnSb films (Fig. 8. 13) the long range crystalline order is very
8.13)
good with quite sharp V/NiMnSb interfaces.

Figure 8. 13 Transmission electron microscope image obtained on a transversal


cut in an NiMnSb (00l)
(001) film grown by MBE at 550 'C on a Vanadium buffer layer
(Turban, 2001).
286 Delia Ristoiu and J. Nozil~res
P. Nozieres
J.-P.

8.4 Magnetism and Transport Properties of NiMnSb


Thin Films

All the magnetic and transport measurements shown in the following refer to
Mo-buffer, FTS-grown films under the best conditions described above
'c and argon pressure 3 x 10- 3 mbar).
(deposition temperature 400 C

8.4.
8. 4. 1 Hysteresis Cycles
The magnetization curves measured with the applied field in and out of the film
plane (Fig. 8.14)
8. 14) clearly show that the magnetization Iies
lies naturally in the film
plane. Considering that the saturation field (0. 9 O. 02) T measured out of
plane corresponds to the demagnetizing field of an infinite film (4nM s ) ' we
can calculate the magnetic moment of NiMnSb. At 10 K we find M s = 3.9
0.21JB per NiMnSb unit formulae. This result is in agreement with the
theoretical predictions (Groot et al., 1983), as well as with the values
obtained on the bulk single crystal (Hordequin et ai.,
al. , 1996, 1997a) .

1.0 """"----------,,,.....---~~
r"....---------:-:".....---~~

0.8

0.6
~vo
,:<'"
~ 0.4

0.2

0.0 <-----'_-----'--_---'--_---L-_-'-----_L----'_
<---'_-'-_-'-_---'-_-'---_L-------'_
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
llo H (T)
lloH(T)

Figure 8. 14 Magnetization curves measured at 10 K for 2 directions of the applied


field: parallel (solid line) and perpendicular (dotted line) to the film surface.

The coercive field lies between 5 and 30 mT, with the larger values for
the thinner films (Fig. 8.15). This behavior is typical of soft magnetic films,
the low thickness enhancement being related to the domain wall pinning at free
surfaces or surface defects.
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 287

30

25

~ 20
l-
f0-
E
;fo 15
~
:::l.
10
'0

o00':---=2'::-0----:'40:----'"60:----:'-SO:------:'-:-00:--------:-!120
0':---::'20:----4-,1:0--6'-::-0------::'8-=-0--:1-!-00:-------,-J120
Thickness (nm)

Figure 8. IS Evolution of the coercive field as a function of the film thickness at room
temperature.

8.
8.4.2
4. 2 Magnetocrystalline Anisotropy
The studies undertaken on bulk NiMnSb single crystals by Hordequin et al.
(Hordequin et al. , 1997b) show that the hard axis is [100J and the easy axis
is [110]. The magnetocrystalline anisotropy, however, is quite small, with a
corresponding anisotropy field of about 25 - 30 mT.
In the case of epitaxial thin films, the situation is reversed with a [100J
[100 J
easy axis and a [11OJ
[110J hard axis. The corresponding anisotropy field is 25 mT
(Fig. 8.16). Magnetoelastic and/or interface contribution to the anisotropy
may be at the origin of such discrepancy. Detailed relaxation studies during
growth will have to be carried out to confirm this point.

1.0
_/-
,
-
/
~

J
O.S
0.8

0.6
~
:i
~~ 0.4
0.2

0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06
prJ!
llo H (T)

Figure 8. 16 Magnetization curves measured at 4 K on a 100 nm NiMnSb film for two in


plane applied field directions: parallel to [100]
[1 OOJ axis (solid line) and parallel to [110J
[110 ]
axis (dotted line).
line) .
288 Nozieres
Delia Ristoiu and J.-P. NoziEHes

Ferromagnetic resonance (FMR) with an in-plane applied field show


symmetrical spectra with Lorentzian shapes (Fig. 8. 17), meaning that the
dispersion phenomena are weak. The resonance linewidth, which is directly
related to the magnetic homogeneity, is also weak, ranging between 10 and
20 mT.

60 80 100 120 140 160 180 200


PaH (mT)
/loH(mT)

Figure 8. 17 Derivative of absorption spectra taken at 300 K.

The angular dependence of the resonance field at room temperature and


10 K are presented in Fig. 8.18.
at 10K 8. 18. The 90'
90 periodic oscillations are a clear
signature of an in-plane four-fold symmetry. In this geometry, the resonance
field H II is given by
2

(W/ ) = H II ( H II + 4nM - 2K
M-II ) (8. 1)

where W II is ferromagnetic resonance frequency in parallel configuration, M is


the magnetization and K II is the in plane anisotropy (Soohoo, 1965; Heinrich
and Bland, 1994). The resonance field is shifted to low (high) fields when the
static field is parallel to an easy (hard) axis. From Fig. 8.18 8. 18 we clearly
[1 OOJ easy direction (e = 45
observe an [100J 45' and 135
135' , minimum of H res re , )) to a

[110J hard direction (e=oo


(8=0' and 90
90' , maximum of H res ) , in agreement with
re , )

the quasi-static measurements.


The anisotropy field can be quantitatively calculated as the half-distance
between the maximum and the minimum of the resonance (Heinrich and Bland,
14 2) mT at 300 K and (202)
1994). Again, the values obtained ((142) (20 2) mT at
10 K) are in good agreement with the quasi-static measurements. We notice a
light increase of the anisotropy at low temperatures.
Concerning the linewidth, the angular evolution follows the resonance
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 289

~ :::1
~ 120
'-;;,
::r::~
:r::~ 110

~
-u
-'5

]
]
20
I!
10
0
_ _...L-_...L-_-'--_-'--_'-----'L-
' : _ - - , - ,_ - - ' - _ - - - - ' - ,_ - , - _ - , - ,_ - . . l ' L -
_

o 30 60 90
!

120
t

150 180
en
Figure 8. 18 Angular dependence of the resonance field and linewidth at 300 K
(solid symbols) and 10K
10 K (open symbols)
symbols)..

field's one (Fig. 8.18), which means that the same physical phenomenon
accounts for the two behaviors. We identified the anisotropy
an isotropy as the source of
resonance field oscillations, consequently it is also the source of the linewidth
oscillations. We can simply explain this relationship considering the magnetic
anisotropy. Under a static field the individual magnetic moments tend to align
on the field direction. Thus if the field is oriented on a hard axis, the moment's
alignment is worse than if the field is oriented on an easy axis. Consequently
we have a magnetic homogeneity variation as a function of the field
orientation, which gives rise to a Iinewidth
linewidth variation. A larger line
Iine corresponds
to the situation when the field is parallel to a hard axis, in agreement with our
observations.

8. 4.3
4. 3 NiMnSb Thin Film Resistivity
Metal resistivity is due to the electron scattering on impurities, on lattice
defects, phonons and magnons, the latter for magnetic materials only. At low
temperature, the phonons and magnons contributions are negligible and the
corresponding so-called residual resistivity is only related to the atomic and
structural disorder of the sample.
The residual resistivity measured on an epitaxial NiMnSb film is (5. 3
0.1)
O. 1) IlQcm.
IJQcm. This value is lower than the resistivity reported on both
polycrystalline thin film, (19 IlQcm
IJQcm (Hordequin et al.,
aI., 1996) and 54 IlQcm
IJQcm
(Moodera et Mootoo, 1994; Cabailero et aI., 1998, and polycrystalline
bulk (6. 9 IlQcm
IJQcm (Otto et al.,
aI., 1987) and 7. 8 IlQcm
IJQcm (Hordequin et al. ,
1996. This is a signature of a very low defects density.
The thickness dependence of the resistivity is plotted in Fig. 8. 19 for a
= 10- 100 nm) /
series of films with the structure NiMnSb (thin film thickness t =
Mo (2 nm). The Mo buffer thickness is chosen thin enough so that its
290 J. P. Nozieres
Delia Ristoiu and J.-P. NoziEHes

contribution is minimal. The film resistance varies linearly with the inverse of
the thickness, clearly indicating that the intrinsic resistivity is thickness
independent in the range considered here. A linear fit yields the slope 4 =~
d( -d
p, where R is the electrical resistance, I = 3 mm represents the distance
between the measure points, L = 5 mm is the sample length and 5 is the
electrical reisstivity. After correction of the Mo buffer resistivity (assuming
resistors in parallel), the deduced average NiMnSb resistivity at room
=(
temperature is p = ( 19.9 1.5) ~Qcm. Again, this value is lower than those of
1. 5) !JOcm.
bulk polycrystals (around 40 - 50 ~Qcm)
!JOcm)..
16

14

12

SID
<l)

g 8
g
if>
'v;
<l)
6
~
4

2
o'--_-'-__-"--_--'-__
O'-----_----'---_ _-'-----_-----'-_ _- '-_---l'--_-'
---'---_-----.J'-----_---..J
0.00 0.02 0.04 0.06 0.08 0.10 0.12
t-IN;MnSb(l/nm)
/-INiMnSb(l/llm)

Figure 8. 19 Thickness dependence of the resistance at 300 K. The continuous line


represents the linear fit of the data points. These measurements were performed on
NiMnSb/ 20 nm Mobilayers in order to reduce the Mo contribution at the resistivity and to
preserve NiMnSb epitaxy.

8.5 Evidences of the Half-Metallic Character

8.5.
8. 5. 1 Overview of Magnetic Excitations in Half-Metals

The half-metallic character is due to the existence of a gap 5 at the Fermi level
in the spin down (minority spins) sub-band. Besides, as for any magnetic
element, an exchange splitting .:1..1 exists between the spin up and spin down
sub-bands (Fig. 8. 20a). For NiMnSb at low temperature, .:1 ..1 and 5 are of the
order of 2 and 1 eV, respectively. The half-metallic gap 5, defined by the
energy difference between the Fermi level and the bottom of the spin down
conduction band, represents only a fraction of the spin-down semiconducting
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 291

band gap.
As the temperature increases, the magnetization decreases due to the
combined effects of a reduction of ..1Ll and of the thermal energy k BB T which
tends to populate states above the Fermi level. Thus the gap 8 {5 may disappear
well before the complete closure of the exchange gap Ll, ..1 , at a temperature T'
at which the Fermi level intersects the bottom of the spin down conduction
band. Hordequin et al. have observed such a crossover in bulk single crystals
r ""'
at T* ~" 100 K. This crossover from a half-metallic to a metallic state was
associated with a transition from a localized
local ized to an itinerant ferromagnetic state
(Hordequin et al. , 2000). The relations between the magnetization and the
density of states in NiMnSb have been theoretically investigated by Kulatov
and Mazin (1990), who predict that the half-metallic character should vanish
for only a weak decrease in magnetization. Irkhin and Katsnelson (1990)
examined the consequences of the spin splitting
spl itting and the influence of spin waves
excitations on the thermodynamic properties, magnetization and relaxation
rate of half-metals. Otto et al. (1987) treated these systems in a strong
coupling model and noticed that" below some temperature T* spin flip
that "below r
scattering is not possible."
possible. "
E E(q)
F(q)
Spin waves
E=D
E=DiJ',
q2
"'
D,(E)

'I
(a) (h)
(b)

Figure 8.20
8. 20 ( a) Schema of the band structure of a half-metal; (b) Schema of the
(a)
magnetic excitations.

The only possible excitations in the half-metallic state are the transverse
coherent spin waves (magnons). Only weak relative deviations of the first
neighbors moments occur and locally all spins, including those of conduction
electrons, remain in the up direction, whereas the long-range correlation
function decreases. In the absence of anisotropy (Goldstone theorem), the
magnetization can rotate as a whole in any direction without any energy
barrier. The conduction electron spins follow the magnetization so that only a
majority spin density still exists at the Fermi level. This rules out the apparent
paradox between the occurrence of a vanishing down spin density and that of
spin excitations.
The Stoner excitations (spin fluctuations) appear at higher temperatures
possib il ity of spin flip
and introduce the possibility fli p scattering. They are incoherent
interband spin-flip excitations that have a more local character and form a
292 Delia Ristoiu and J. -P. Nozif~res
J.-P. Nozil~res

continuum as a function of the energy (Gautier, 1982). In the case of half-


metals, their minimum energy correspond to the distance 8 {j between the Fermi
level and the bottom of the spin down conduction band (Fig. 8. 20b). This
minimum energy of the Stoner continuum occurs for a wave vector defined by
the difference between the Fermi vector of the majority band and the
reciprocal lattice vector corresponding to the bottom of the minority band (the
X point of the Brillouin zone in the case of NiMnSb).

8. 5. 2
8.5.2 Electrical Resistivity

Starting with the studies of Otto et al. (1987) on bulk NiMnSb, an anomaly in
the resistivity behavior was evidenced around T' = 100 K. This type of
discontinuity was later explained as a crossover from a low temperature half-
metallic state to a normal metallic state above rT' (Hordequin et al., 1996).
This anomaly has been also observed in thin films (Moodera and Mootoo,
1994; Cabailero et aI.,
al., 1998; Hordequin et al., aI., 1997a) and recently in
epitaxial films as can be seen in Fig. 8.21
8. 21 for a thick 100 nm NiMnSb with a
thin 2 nm Mo buffer layer in order to minimize the Mo contribution to the
resitivity.
30 /
T2/
/
/
/ r l65
20 /
/
Eu /
a
c: /
.3
2: /.i
Q 10

o 50 100 150 200 250 300


Temperature (K)

Figure 8. 21 Thermal variation of the resistivity; continuous line represents the T 165
8.21 IG5
fit
for T> T' and the dotted line represents the T fit for T < T' . The samples are 100 nm
2

NiMnSb/ 2 nm Mobilayers in order to preserve NiMnSb epitaxy (Mo conditioned) and to


have a minimum Mo contribution on the resitivity (thick NiMnSb, very thin Mo).
Mo) .

In order to explain the nature of this anomaly we proceeded to a fit of the


two regions, which results are as follows:
follows;
P (T) = POI + aT 2 , for temperatures below T * (8.2)
= P02
( T) =
P (T) + bTl 65, for temperatures above T'
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 293

where POI = = 5.3 ~Qcm, a =


5. 3 !JOcm, = 5.8 xX 10- ~Qcm/K2,
1O- 4 !JOcm/K 2
= 6. 1 !JO
, P02 = =
~Q cm, b =
1.7 1O-3~Qcm/K165.
1. 7 1O-3!JOcm/KI65.
The electrical resistivity P of a non-magnetic material is mainly due to
defects and phonons scattering. In a ferromagnetic material an additional
contribution due to the interaction between the magnetic moments of the
conduction electrons and the magnetic excitations (spin waves and/ or Stoner
fluctuations) exists. Scattering by an existing spin wave (temperature excited
spin wave, for example) preserves the conduction electron spin, for the spin
wave energy is lower than a single spin flip energy. On the other hand,
scattering by a Stoner spin fluctuation allows spin flip. At low temperatures
both contributions vary like T 2 (Mills and Lederer, 1966), while at higher
temperatures spin waves give a T contribution and the Stoner spin fluctuations
give a T 5/ 3 contribution CUeda and Moriya, 1975).
The fact that the low temperature resistivity is lower than predicted by the
high temperature fit can be interpreted as resulting from the extinction of a
scattering mechanism. Thus if T' T * marks the transition to a half-metall ic state,
the spin flip scattering are forbidden below this temperature (no possible final
states) and only the spin waves and the phonons can contribute to the
resistivity. The low temperature fit corresponds indeed to a T 2 variation that
represents spin waves only contribution (the phonon term is small in this
temperature range). Above T' T * we may have a superposition of different
contributions: spin waves, Stoner spin fluctuations and phonons.
Consequently, the exponent {3, f3, characterizing the thermal behavior of the
resistivity, has to be in an intermediate value, between that of phonons, spin
waves and spin fluctuations. The value of {3 f3 = 11.65
.65 is close to the Stoner spin
fluctuations exponent (f3 = 1.66)
({3 = 1. 66) and slightly higher than that obtained for the
bulk single crystal ({3
(f3 = 1. 35, Hordequin et al. , 1996).

8.5.3 Magnetoresistance
Magnetoresistance (MR) characterizes the transport properties of a material in
an external magnetic field. In a normal ferromagnet in the saturated state (no
domains), we expect two contributions to the magnetoresistance: cyclotron
MR, which is observed in any metal, and characterized by a parabolic
increase of the resistance at large fields, and spin disorder MR, which is
characterized by a decrease of the resistance with increasing field. The former
is due to the Lorentz force acting on the moving electrons, while the latter is
due to the progressive spin al ignment within the ferromagnet. The cyclotron
alignment
contribution to the MR decreases with the temperature as the electron mean
free path decreases, while the spin disorder contribution to the MR should
increase with temperature. For a multidomain sample, an additional low field
contribution of same sign as spin disorder MR is also observed, which
corresponds to the progressive alignment of the different domains along the
294 Delia Ristoiu and J.-P. Nozieres
Nozil!res

applied field.
In a half-metal ferromagnet, all spins are supposed to be along a single
direction, so that no spin disorder magnetoresistance should be observed.
However, if we consider that spin waves may exist in the half-metallic state, a
little spin disorder contribution to the MR, due to the misalignment of the spins
around the equilibrium spin up direction, is also expected.
Both cyclotron and spin disorder MR are observed in epitaxial NiMnSb
films depending on temperature (Fig. 8.22). At low temperature, the MR is
clearly of cyclotron type (positive slope). The small low field contribution
(negative slope) corresponds to the domain al ignment within the film, which is
multidomain in the remanent state. At room temperature, the MR exhibits a
negative slope characteristic of spin disorder MR. By plotting the high field MR
slope (from 2 to 4 T) as a function of the temperature (Fig. 8. 23) we
=
evidence a change in sign around T' = 110 K. This discontinuity point
corresponds to the same temperature as for the resistivity thermal behavior
change. It is coherent with the picture of a closure of the half-metall
half-metallic
ic gap 0,
(j,
which allows the onset of spin disorder magnetoresistance.

-+- 5K

~ 0.1

~
o 2 4
lloH (T)

~ -0.2

~
-0.4

o 2 4
lloH (T)

Figure 8. 22 Magnetoresistance curves measured at 5 and 300 K.

8. S.
5. 4 Hall Effect

For a non-magnetic material, the Hall voltage varies linearly with the field, but
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 295

0.\ 5
0.15

~ 0.\0
0.10
t::
t:::
e
~
~ 0.05
:::l.
~
~
o - - - -- - - - - - -
~

~ -0.05 I I
:u
:0 I ! I!
-0.\0
-0.10
!I
II
- 0.\ 5 '-_--L.._----'-_
-0.15 '--_--'-_-----'-_ _L--_--L.._-----'-_------'
_'----_--'-_---'-_----..J
o SO 100 1SO
\ SO 200 250 300
Temperature (K)

Figure 8.23 Thermal variation of the magnetoresistance slope. Both magnetic field and
current are applied on the [110]
[110J NiMnSb crystal direction.

for a magnetic material an anomalous Hall voltage, proportional to the


magnetization, is superimposed. The magnetic field intensity Hand
p Xy can be written as (Hurd, 1972):
temperature T dependent Hall resistivity PXY
(8.3)

where R H ( T) and R A ( T) are respectively the normal and anomalous Hall


coefficients.
Hall effect measurements are generally performed on patterned samples in
order to have well defined current lines (see insert of Fig. 8.24b).
8. 24b). A typical
Hall resistance curve, measured at 10K 10 K on an epitaxial NiMnSb film, is
presented in Fig. 8. 24a. The saturation at 0.9 T corresponds to the full out of
plane rotation of the magnetization (Fig. 8. 24a). At saturation R A becomes
110 RAMs, where M
1-10 Mss is the saturation magnetization at the measured
temperature. Above the saturation field, the only contribution is that of R HH
which presents, as expected, a Iinear
linear behavior. For the free electron model
p Xy above
(Hurd, 1972), the normal Hall coefficient deduced from the slope of PXY

0.9 T is directly related to the carriers density n: R H ~t


= nq where B is the

magnetic induction, q == 1.
1.66 x
X 10
19
10--19 C the elementary charge and t the sample
thickness (100 nm here). R H is positive over all the temperature range studied
here, i. e. from 5 to 300 K. This means that the current carriers in NiMnSb are
holes. The carrier sign can be explained by the fact that the conduction band
closest to the Fermi level has a hole character (Hordequin et al. , 2000).
The temperature dependence of the number of carriers per NiMnSb unit
formulae is reported in Fig. 8. 24b. Again, an anomaly around T' = 100 K is
observed. The carriers number above T* T' , e. g. 4 holes per NiMnSb unit
formulae, is quite constant, whereas it decreases below T' . This decrease of
the Hall coefficient below T' cannot be described in the framework of the
296 Nozh~res
Delia Ristoiu and J.-P. NoziEHes

0.010


0.008

0.006 J1 0 RsM
llo
:
>..

c< 0.004

0.002

o 0.5 1.0 1.5 2.0 2.5 3.0 3.5


J1a
lloHH (T)
(a)
5

1il~
"3 3
:;

o
u...
11~ 2
"0
:r:
J:

o 50 100 150 200 250 300


Temperature (K)
(b)

Figure 8. 24 (a) Hall resistance as a function of field with schematic illustration of the
different contributions; (b) Thermal variation of the carrier's number per unit formulae.

classical Hall effect model, according to which the Hall coefficient is


temperature independent. Nevertheless, this variation may be understood
considering that the mobility of the carriers has to increase when crossing at
T* from a normal metallic state to a half-metallic state, because the spin flip
scattering is no longer possible.
Note that the minority electron pocket around the X point of the Brillouin
zone is shifted above the Fermi level and becomes unoccupied at T' , thus
reducing the electron character of the transport properties. However, there
are still three bands of majority spins (spin up) crossing the Fermi level
(Hanssen and Mijnarends, 1986), one of electron character and two
corresponding to open surfaces (holes). This prevents any precise calculation
of an effective density of carriers or mobility from the Hall effect and resistivity
data, as previously pointed out by Otto et al. (1989a, 1989b).
1989b) .
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 297

8. S. 5
8.5. Magnetization

For an itinerant ferromagnetic material the magnetization decrease with the


increase of the temperature is mainly due to the Stoner spin fluctuations and

= M~ [ 1- b ( ~ ) 2] (Mook, 1988). In half-metals


corresponds to the law: M 2=

the Stoner spin fluctuation is forbidden and the magnetization decrease with
temperature can be explained only by spin wave excitation whose thermal
behavior follows a T 312
3 2
/ law: M == M s (1 - AT 3312
( 1-
2
/ ) .

The temperature dependence of the magnetization measured in the film


plane under a field of 0.2 T is reported in Fig. 8. 25a. As we can see on our
plot representing M 2 as a function of T 2 , this law is well verified in the high
temperature regime.
A deviation with respect to the T 2 law appear around T' = 100 K. Below

16.5

16.0

~
N
"'co
.3
.3 15.5
N
'"
~
15.0

14.5
0 I x 104 2x 104 3x 104 4x 104 5x 104
rT22 (K2)
(a)
16.2

16.1

~
~
co
3::l.
~
~ 15.9

15.8

15.7
0 200 400 600 800
TJI2 (K312)
rJl2
(b)

Figure 8.
8.25 M2 (T 2
25 Temperature dependence of the magnetization (a) M ) from 10 to
3 2
250 K; (b) M(T / ) from 10 to 100 K.
M(T3/2)
298 Nozh~res
Delia Ristoiu and J.-P. Nozieres

T' the measured magnetization is larger than predicted by M 2 ( T 2 ) law. A M


( T 3 / 2 ) fit for T < T' is reported in Fig. 8. 25b. The agreement is very good
which clearly suggests that only spin wave excitations are present. The
vanishing of the spin fluctuations below T' = 100 K may be considered as the
signature of a low temperature half-metallic
half-metal Iic state. Inelastic neutron diffraction
measurements undertaken by C. Hordequin et al. (1997b) suggest that this
transition may be equally related to the evolution of the system from a
localized magnetism below T T'* to an itinerant magnetism above T * .
It is interesting to note that at the crossover temperature, the
magnetization is only slightly sl ightly reduced from its zero Kelvin value:
M ( rT * )/ =
) / M (0) = 99.5 %. The mean band spl itting .t1 ..1 between the spin up and
spin down sub-bands should present a variation proportional to the
magnetization, so it has to be just sl ightly reduced at the crossover
temperature.

8.6 Conclusions

Epitaxial NiMnSb thin films can be prepared by several techniques, with


structural and physical properties equivalent if not better than those of bulk
single crystals. In particular, a clear evidence of a transition from the low
temperature half metallic state to a normal state is observed around a critical
temperature close to 100 K. Recent photoemission experiments have also
revealed that a well prepared, clean, ordered and stoichiometric surface
exhibits a large spin polarization, as expected for half metals (R istoiu et al. ,
2000a). Furthermore, the complete heteroepitaxy of a NiMnSb/MgO/ (V)
NiMnSb trilayer has been recently demonstrated (Turban, 2001) as well as the
homoepitaxy of NiMnSb with the semiconducting TiCoSb Heusler compound
(Ristoiu,
(R istoiu, 2000d). All this opens up the door for a further integration of these
films into more complex structures such as GMR, TMR and hybrid
semiconductor and superconductor / ferromagnet multilayers. Ultimately, fully
functional devices will be developed, in which the full spin polarization of
NiMnSb will be used to generate enhanced signals and/or new functions. The
era of spin electronics is only at a burgeoning state, and there is no doubt that
half metals will playa key role in its development.

References
Attaran, E. and P.J. Grundy. J. Magn. Magn. Mater. 78: 51 (1989)
Baibich, N. M. , J. M. Broto, A. Fert, Van Dau, N. Petroff, P. Etienne, G.
Creuzet, A. Frederick and J. Chazelas. Phys. Rev. Lett. 61: 2472 (1988)
FrederickandJ.
Bauer, E. and J.H.
J. H. van des Merwe. Phys. Rev. B 33: 3657 (1986)
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 299

Bobo, J. F. , K. Bessho, F. B. Mancoff, P. R. Johnson, M. C. Kautzky, R. L.


White and B. M. Clemens. Mat. Res. Soc. Symp. Proc. 475: 21 (1997a)
Bobo, J. F. , P. R. Johnson, M. C. Kautzky, F. B. Mancoff, E. Tuncel, R. L.
White and B.M.
B. M. Clemens. J. Appl. Phys. 81: 4164 (1997b)
Bona, G. L.L.,, F. Meier, M. Taborelli, E. Bucher and P. H. Schmidt. Solid
State Comm. 56: 391 (1985)
Bozorth, R. M. In: Ferromagnetism. D. van Nostrand Co, New-York p. 328
( 1951)
1951>
Caballero, J.A.,
J. A. , Y. D. Park, A. Cabbibo, J.
Y.D. J.R.R. Childress, F. PetroffandR.
Petroff and R.
Morel. J. Appl. Phys. 81: 2740 (1997>
Caballero, J. A. , Y. D. Park, J. R. Childress, J. Bass, W. C. Chiang, A. C.
Reilly, W. P. Pratt jr. and F. Petroff. J. Vac. Sci. Technol. A 16: 1801
( 1998a)
Caballero, J. A. , W.J.
J.A., W. J. Geerts, F. Petroff, J.U.
J. U. Thiele, D. WelierandJ.R.
Weller and J. R.
Childress. J. Magn. Magn. Mater. 177-181: 1229 (1998b)
Caballero, J.A.,
J. A. , A.C.
A. C. Reilly, Y. Hao, J. Bass, W. W.P.
P. Prattjr.,
Pratt jr. , F. Petroff
and J.R. Childress J. Magn. Magn. Mater. 198-199: 55 (1999)
Castell iz, Monatsh. Chem. 82: 1059 (1951> (1951)
Childress, J. R. , J. A. Caballero, W. J. Geerts, F. Petroff, P. Galtier, Y.
Suzuki, J.U.
J. U. Thiele and D. Weller. Mat. Res. Soc. Symp.Proc. 475: 15
((1997)
1997>
Dieny, B., V.S. Speriosu, S.S.P. Parkin, B.A. Gurney, D.R. Wilhoit and
D. Mauri. Phys. Rev. B 43: 1297 (1991> (1991)
Endo, K. J. Phys. Soc. Jpn. 29: 643 (1970)
Engen, P. G. van, KH. J. Bushow, R. Jongebreur and M. Erman. Appl.
Phys. Lett. 42: 202 (1983)
Fujii, S. , S. Sugimura, S. Ishida and S. Asano. J. Phys. : Condens. Matter
43: 8583 (1990)
Gautier, F. In: Magnetism of Metals and AI/oys.Alloys. North Holland, Amsterdam
(1982)
( 1982)
Groot, R. A. de, F. M. Mueller, P. G. van Engen and K. H. J. Buschow, Phys.
Rev. Lett. 50: 2024 (1983)
Groot, R.R.A.de
A. de and K.H.J.
K. H. J. Bushow. J. Magn. Magn. Mater. 54-57: 1377
(1986)
( 1986)
Groot, R.A.de. Physica B 172: 45 (1991> (1991)
Hamzic, A.,A. , R. Asomoza and I. A. Cambell. J. Phys. F: Metal Phys. 11:
1441 (1981>
(1981)
Hanssen, K.E.
K. E. and P.E.,
P. E. , Mijnarends. Phys. Rev. B B34:
34: 5009 (1986)
Hanssen, K. E. , P. E. Mijnarends, L. P. Rabou and K. H. J. Bushow. Phys.
Rev. B 42: 1533 (1990)
Heinrich, B. and J. A. C. Bland. In: Ultrathin Magnetic Structures IT.
Springer-Verlag, Berlin, Heidelberg (1994)
Helmholdt, R. B. , R. A. de Groot, F. M. Mueller, P. G. van Engen and K. H.
J. Bushow. J. Magn. Magn. Mater. 43: 24 (1984)
300 Delia Ristoiu and J.-P. Nozieres
NoziEHes

Heusler, F. Verhandl. Deut. Physik. Ges. 219 (1903)


Hordequin, C., C. , J. Pierre and R. Currat. J. Magn. Magn. Mater. 162: 75
(1996)
Hordequin, C., C, E. Lelievre-Berna and J. Pierre. Physica B 234-236: 603
( 1997a)
Hordequin, C. Thesis. J. Fourier University Grenoble- France (1997b)
Hordequin, C. , J. P. Nozieres and J. Pierre. J. Magn. Magn. Mater. 183:
225 (1998)
Hordequin, C. , D. Ristoiu, L. Ranno and J. Pierre. Euro. J. of Phys. B 16:
287 (2000)
Hurd, C. M. In: The Hall Effect in Metals and Alloys. New-York Plenum
( 1972)
Inukai, T. , M. Matusuoka and K. Ono. Appl. Phys. Lett. 49: 52 (1986)
Irkhin, VVY. andM.I.
Y. and Katsnelson. J.
M.1. Katsnelson J.Phys.:
Phys. : Condo Matter 2: 7151 (1990)
Ishida, S., S. S Kashiwagi, S. Fujii and S. Asano. Physica B 210: 140
(1995a)
Ishida, S. , S. Fujii, S. Kashiwagi and S. Asano. J. Phys. Soc. Jap. 64:
2152 (1995b)
Ishida, S., T. Masaki, S. Fujii and S. Asano. Physica B 239: 163 (1997)
Jull iere, M. Phys. Lett. 54: 225 (1975)
Julliere,
Kabani, R., M. Terada, A. Roshko and J.S. Moodera. J. Appl. Phys. 67:
4898 (1990)
Kaczmarska, K., K. , J. Pierre, J. Beille, J. Tobola, R.V.
R. V. SkolozdraandG.A.
Skolozdra and G. A.
Melnik. J. Magn. Magn. Mater. 187: 210 (1998)
Kaczmarska, K. , J. Pierre, J. Tobola and R. V. Skolozdra. Phys. Rev. B
60: 373 (1999)
Kautzky, M. C. and B. M. Clemens. Appl. Phys. Lett. 66: 1279 (1995)
Kautzky, M. C C. , F. B. Mancoff, J. F. Bobo, P. R. Johnson, R. L. White and
B. M. Clemens. J. Appl. Phys. 81: 4026 (1997>
B.M.
Kouacou, M. A. , J. Pierre and R. V. Skolozdra. J. Phys. Condens Condens.. Matter
7: 7373 (1995)
Kubler, J. In: Proc. Workshop on 3d 3 d Metallic Magnetism. ILL-Grenoble,
private communications, p.99p. 99 (1983)
Kulatov, E. and I. I. Mazin. J. Phys. : Cond. Matter 2: 343 (1990)
Lewis, B. and J. C. Anderson.
Anderson In: Nucleation and Growth of Thin Films.
Academic Press (1978)
Mancoff, F. B. , B. M. Clemens, E. J. Singley and D. N. Basov. Phys. Rev.
R 12 ,565 (1999)
B 60: R12,565
Mills, D.L.
D. L. and P. Lederer. J. Phys. Chem. Solids 27: 1805 (1966)
Moodera, J.S. and D.M. Mootoo. J. Appl. Phys. 76: 6101 (1994)
Moodera, J.S., G. J. Mathon, O. Magn. Magn. Mater. 200: 248 (1999)
Mook, H. A. In: Spin Waves and Magnetic Excitations. by A. S. Borovik-
Romanov and S.K.Sinha
S. K. Sinha eds. Elsevier Science Publ. p.425
p. 425 (1988)
Moran-Lopez, J. L., R. Rodriguez-Alba and F. Aguila-Granja. J. Magn.
Monocrystalline Half-Metallic NiMnSb Thin Films: Preparation... 301

Magn. Mater. 131: 417 (1994)


Naoe, M. , N. Kitamura, M. Shoji and A. Nagai. J. Appl. Phys. 63: 3636
(1988)
Naoe, M. , N. Matsushita and S. Nakagawa. J. Magn. Magn. Mater. 134:
395 (1994)
Ohyama, R. , J. Abe and K. Matsubara. IEEE Trans. J. Magetics in Japan 1 :
122 (1985)
Ohyama, R., T. Koyanagi and K. Matsubara. J. Appl. Phys. 61: 2347
( 1987)
D . H. Fujiwara, T.
Orgassa, D., C. SchulthessandW.H.
T.C. Schulthess and W. H. Butler. Phys. Rev. B
60: 13,237 (1999)
Otto, M.J.,
M. J. , H. Feil, R.A.
R. A. van Woerden, J. Wijngaard, P.J.P. J. van der Valk.
Valk,
C. F. van Bruggen and C. Haas. J. Magn. Magn. Mater. 70: 33 (1987)
Otto.
Otto, M.J., R.A. vanWoerden, P.J. vanderValk, J. Wijngaard. Wijngaard, C.F. van
Bruggen, C. Haas and K. H. J. Buschow. J. Phys. Cond. Matter 1: 2341
(1989a)
Otto, M.J.,
M.J . R.A. vanWoerden,
van Woerden, P.J. van der Valk,Valk. J. Wijngaard, C.F. van
Bruggen and C. Haas. J. Phys. Condo Matter 1: 2351 (1989b)
Parkin, S.S.P. Phys. Rev. Lett. 71: 1641 (1993)
Pickett, W.E. and D.J. Singh. Ph'ys. PhYs. Rev. B 53: 1146 (1996)
Pierre.
Pierre, J. , R. V. Skolozdra.
Skolozdra, J. Tobola, S. Kaprzyk, C. Hordequin.
Hordequin, M. A.
Koucou.
Koucou, I. Karla, R. Currat and E. Lei ievre-Berna. J. of Alloys and
Compounds. 262-263:
262 - 263: 101 (1997)
Ristoiu D. , J. P. Nozieres, C. N. Borca, T. Komesu, H-K. Jeong and P. A.
Dowben. Europhys. Lett. 49. 624 (2000a)
Ristoiu, D. , J. P. Nozieres, C. N. Borca.Borca, B. Borca and P. A. Dowben.
Appl. Phys. Lett. 76: 2349 (2000b)
Ristoiu, D.,
D. , J.P.
J. P. Nozieres and L. Ranno. J. Magn. Magn. Mater. 219: 97
(2000c)
Ristoiu, D. Thesis. J. Fourier University Grenoble -France (2000d)
Robey, S. W., L. T. Hudson and R. L. Kurtz. Phys. Rev. B 46: 11,697
(1992)
Rodriguez-Alba, R. R.,, F. Aguila-Granja and J. L. Moran-Lopez. In: New trends
Aguila-GranjaandJ.L.
in Magnetism: Magnetic Materiales and their Applications. J. L. Moran-
Lopez ed. Plenum Press,
Press , New-York, p. 299 (1994)
van Roy, W. , J. de Boeck, B. Brijs and G. Borghs. Appl. Phys. Lett. 77:
4190 (2000)
Schlomka, J. P., M. Tolan and W. Press. App!. Appl. Phys. Lett. 76: 2005
(2000)
Schwarz, K. J. Phys. F Metal Phys. 16: L211 (1986)
Shiomi. S. T. Maegawa.
Shiomi, S., Maegawa, H. Iwakiri and M. Masuda. Jap. J. of Appl.
Phys. 27: L1718 (1988)
Shoji. M . A. Nagai.
Shoji, M., Nagai, N. Murayama, Y. Obi and H. Fujimori. IEEE Trans. J.
Magetics in Japan, 2, 698 (1987)
302 Delia Ristoiu and J. -P. Nozit~res
NoziEHes

Soohoo, R. F. In: Magnetic Thin Films. Harper & & Row Inc. (1965)
Soulen Jr. , R. J. , J. M. Byers, M. S. Osofsky, B. Nadgorny, T. Ambrose,
S. F. Cheng, P. R. Broussard, C. T. Tanaka, J. Nowak, J. S. Moodera,
J. M. D. Coey. Science 282: 85 (1998)
A. Barry and J.M.D.
Soulen Jr., R. J., M. S. Osofsky, B. Nadgorny, T. Ambrose, P. R.
Broussard and S. F. Cheng. J. Appl. Phys. 85: 4589 (1999)
K. , H. Fujimori, M. Shoji and A. Nagai. Jap. J. of App!.
Takanashi, K., Appl. Phys.
26: L1317 (1987)
Tanaka, C. T., J. Nowak and J. S. Moodera. J. App!. Appl. Phys. 81: 5515
( 1997)
Tanaka, C. T., J. Nowak and J. S. Moodera. J. Appl. Phys. 86: 6239
(1999)
(1999 )
Thornton, J. A. In: Coating Deposition by Sputtering, Deposition
Technologies for Films and Coatings. Noyes Publications, Park Ridge, New
Jersey (1982)
Tobola, J. , J. Pierre, S. Kaprzyk, R. V. Skolozdra and M. A. Koucou. J.
Phys. Condens. Matter 10: 1013 (1998)
Turban, P. Thesis. H. Poincar University Nancy-France (2001)
Ueda, K. and T. Moriya. J. Phys. Soc. Japan 39: 605 (1975)
Yanase, A. and K. Siratori. J. Phys. Soc. Jpn. 53: 312 (1984)
Youn, S.J. andB.1.
andB.I. Min. Phys. Rev. 851:B51: 10,436 (1995)
Webster, P. J. and R. M. Mankikar. J. Magn. Magn. Mater. 42: 300 (1984)
9 Bulk Amorphous Magnetic Materials

H. Chiriac and N. Lupu

9. 1 Introduction

The interest in glass-forming metallic alloys is coming back nowadays due to


the requirements of the advanced technological applications as well as to the
necessity to understand the mechanisms governing their macroscopic
properties in order to avoid the malfunctioning of the equipments using them.
There are also strong requirements for the development of new amorphous
materials in well-defined shapes.
Metallic glasses can be synthesized in a rich variety of systems and
shapes at cooling rates of 10 5 - 106 K/s: thin films (Cahn, 1993), melt-spun
ribbons with thicknesses up to 50 IJm (Shingu and Ishihara, 1993), wires with
diameters up to 150 - 200 IJm (Masumoto et al., aI., 1985), glass-covered wires
with the metallic core diameter below 50 IJm (Wiesner and Schneider, 1974),
powders (Schwarz, 1993), and more recently bulk samples (Inoue et al. ,
1993). Their interesting magnetic properties in the as-quenched state,
combined with good mechanical properties and high corrosion resistance make
them very attractive for applications and one viable concurrent for the
crystalline materials. In the last two decades, amorphous alloys were used
widely as precursors for nanocrystalline (Yoshizawa et ai., aI., 1988) or
quasicrystalline (Inoue and Kimura, 1999) alloys, which exhibit remarkable
magnetic and mechanical properties due to their very complex structure.
In spite of that the amorphous alloys as ribbons, sheets, wires, glass-
covered wires, powders or thin films are used nowadays in different
engineering applications (cores transformers, magnetic shields, sensors and
transducers, solders or brazing elements, catalysts, etc.). The interest in
glass-forming alloys, which vitrify at relatively low cooling rates (10 - 102 K/s)
from the molten state compared with conventional rapidly quenched metallic
glasses, has grown in the last years. Owing to their resistance to
crystallization, these easy glass-forming alloys can be cast in bulk shapes with
dimensions of millimeters.
The development of these new amorphous alloys contributed widely to the
understanding of the behavior of the supercooled Iiquid liquid regions before
amorphization, due to the high thermal stability of their wide supercooled
304 H. Chiriac and N. Lupu

liquid regions, making also possible the design of new applications based on
their ability to be prepared in three-dimensional shapes.
The history of bulk metallic glasses dates back in time. The first results on
bulk metallic
metall ic glasses were reported in 1974 by the researchers from Bell
Laboratories (Chen, 1974). They succeeded in the preparation of cylinders cyl inders
with diameters up to 3 mm in Pd-Cu-Si system at cooling rates no larger than
10 3 K/
K/s.
s. Afterward, Turnbull and his co-workers prepared and studied many
other bulk amorphous samples with diameters up to 10 mm in Pd-Ni-P ternary
system (Drehman et al., 1982; Kui et al., 1984, 1985). They prepared
these samples at cooling rates of few K/s by quenching the molten alloy in an
B2 0 3 atmosphere. But due to the high cost of Pd these studies were
considered at that time just as simple"
simple "laboratory
laboratory curiosities. "
A few years later, 2 groups of researchers, one from Institute for
Materials Research, Tohoku University, Sendai, Japan and the other one from
University of Virginia, USA, started to independently study the possibility to
prepare bulk amorphous alloys in AI-based multicomponent systems (Inoue et
al. , 1988; He et al., 1988), due to the new requirements of the industrial
applications (especially aeronautic industry), for new materials with low
density and high performances (good mechanical properties combined with
high mechanical strength and good ductility).
These studies were extended immediately to many other systems, and the'
possibility to prepare bulk amorphous samples in different pre-defined shapes
became a reality. In 1989, Inoue et al. from Tohoku University, Japan,
started to study and to develop new multicomponent alloys with easy fusible
eutectic compositions (for which liquid phase is stable at low temperatures) ,
capable to be prepared in bulk shapes with dimensions from a few millimeters
to centimeters by conventional casting methods. Starting with La-based and
Mg-based quaternary alloys, Inoue and his co-workers extended the range of
large glass forming ability systems to Zr-, Fe- and Ni-based alloys. An
overview on the study of these systems, including glass-forming ability and
physical properties, was published
publ ished in 1998 (Inoue, 1998). In almost the same
period, Johnson et al. from California Institute of Technology, USA,
developed and studied a new family of large glass-forming alloys based on
Zr-Ti and containing additions of Be (but no more than 25 % ). These alloys are
better known under their trademark, VITRELOY, and usually require
amorphization cool ing rates of about 1 K/
cooling K/ss (Peker and Johnson, 1993). They
can be obtained as completely amorphous bars or plates with thicknesses of
5 - 10 mm and, due to their special mechanical performances, are used
especially in the sports goods industry (golf clubs or tennis) .
Nowadays, bulk amorphous alloys are proposed, and one of them is used
successfully, for a wide range of applications including micro-gears, welding
elements, dental and other medical implants or prosthesis, writing goods,
sport goods, casting shapes, solders and brazing elements, etc.
All these bulk amorphous alloys are not ferromagnetic at room
Bulk Amorphous Magnetic Materials 305

temperature. There were many efforts to accomplish this goal either by the
increase of the amount of the transition metal in the above-mentioned large
glass-forming systems or by the addition of new ferromagnetic transition
metals. However.
However, the only result was the decrease of the glass-forming abil ity
ability
and the impossibility to prepare amorphous samples in bulk shapes.
No results on bulk amorphous alloys with ferromagnetic properties at room
temperature were reported before 1995. Inoue et al. from Tohoku University, University.
Sendai, Japan announced in 1995 the preparation of the first bulk amorphous
Sendai.
alloys with soft magnetic properties at room temperature in a Fe-( AI. AI, Ga)-
(P.C.B.Si)
(P,C,B,Si) multicomponent system (Inoue and Gook. Gook, 1995a,1995a. 1995b; Inoue
et ai.,
al . 1995c),
1995c). and one year later they announced the first bulk amorphous
hard magnetic alloys synthesized in (Nd. (Nd, Pr) -Fe-AI system (Inoue et al. ,
1996a, 1996b). These alloys are very attractive for the subsequent development of
1996a.
basic science and engineering applications.
applications, but so far the mechanisms
governing their magnetic behavior are not well known and understood.
Three empirical rules responsible for the enlargement of the glass-forming
ability were proposed (Inoue.
(Inoue, 1995):
( 1) the system should contain at least 3 component elements;
(2) significantly different atomic size ratios between main elements above
about 12%;
12 % ;
(3) negative heats of mixing between components.
These rules were defined empirically after many experimental
observations of large glass-forming ability systems.
systems, but they can be explained
in agreement with the kinetic and thermodynamic theories of the solids.
This chapter is restricted to the magnetic properties of two families of
bulk amorphous alloys that we consider representative by far for this field: soft
magnetic (Fe. Co. Nj)
(Fe, Co, Ni) -Zr-B based alloys and room temperature high-coercivity
Nd-Fe-(AI, Si) alloys. Emphasis is placed on the interplay between the atomic
Nd-Fe-(AI.
structure and the magnetic properties. The first part analyzes the techniques
used for the processing of bulk amorphous materials. Particularities both about
the single roller melt spinning method used to prepare thick ribbons with
thicknesses up to 200 IJm and the mold casting or suction casting techniques
used to prepare bulk samples like rods or rings are presented in detail. The
techniques and equipments used for structural and magnetic characterization of
bulk amorphous samples are briefly presented.
composition, preparation
The second part discusses the influence of composition.
technique (cooling rate).
rate), annealing and the on-set of crystallization on soft
magnetic properties of (Fe,Co,Ni)7o(Zr,Nb,M),oB
(Fe. Co. Ni )70 (Zr. Nb. M) 10 B2o =
20 (M=Zr,
(M Zr. Ti. Ti, Ta.
Ta, Mo)
bulk amorphous alloys. The influence of composition on the glass-forming
ability is presented in detail. The evolution of the magnetic characteristics
(saturation magnetization, field, effective magnetic permeability,
magnetization. coercive field. permeability.
saturation magnetostriction) for amorphous ribbons and bulk amorphous rods
and rings is presented comparatively.
Ndgo - x Fe x A1 1o - ySi y(x = 35 at. %
The third part deals with high-coercivity Ndgo-xFexAllo-ySiy(x=35
306 H. Chiriac and N. Lupu

= 0 at. % - 10 at. %) bulk amorphous alloys, presenting at the


- 50 at. %; Y =
beginning the structural measurements by means of X-ray diffraction and
neutron diffraction measurements. Some data about the formation of the
amorphous phase obtained through calorimetric measurements are presented.
Then, dc- and ac-magnetic measurements from low to high fields and
temperatures ranging from 4 to 1000 K are presented in detail. Finally,
M6ssbauer spectroscopy results are listed, pointing out the differences in the
Mossbauer
Fe atoms environments as a function of the addition element, AI or Si.
The last part analyzes the relationship between microstructure and
magnetic properties in bulk amorphous alloys. Brief remarks conclude the
chapter. Finally, some potential applications are suggested.

9.2 Processing and Characterization of Bulk Amorphous


Magnetic Alloys

In the following, we presume (in agreement with other authors) the melt-spun
ribbons as bulk samples for two reasons: the preparation technique is a
continuous casting method and the dimensions of the ribbons, including
thickness, are higher than the dimensions of evaporated or deposited
amorphous thin films.
For the processing of bulk amorphous alloys, master alloys of 5 - 109
each, were prepared by arc melting pure elements (99. 99 %) in an argon
atmosphere, and re-melted several times for homogenization.

9.2.1 Melt-Spinning Technique Used for the Preparation of "Bulk"


Amorphous Ribbons
The usual and very well known method for the preparation of metallic glasses
like ribbons, especially amorphous magnetic ribbons, is the single roller melt
Iike
spinning technique. There are many reports on this technique from the first
evidence for the preparation of the amorphous phase in Au-Si binary system by
Duwez and co-workers in 1960 (K lement et ai.,
(Klement aI., 1960). The method consists
of the ejection of the molten alloy on a rotating support by applying an ejection
pressure. The cooling rates assured in this method range from 10 4 to 106 K/s,
enough for amorphization. Melt-spinning ribbons are amorphous if two basic
conditions are satisfied:
(1) the ribbon's thickness is small enough, so the sol idification cooling
solidification cool ing
rate is higher than the critical cooling rate required for amorphization;
(2) the ribbon remains in contact with the surface of the rotating wheel
long enough, so its temperature is smaller than the crystallization temperature
Bulk Amorphous Magnetic Materials 307

( T x), when it leaves the surface of the wheel.


The fulfillment of these 2 conditions depends on the correlation between
the preparation parameters: wheel velocity, ejection pressure, shape and
size of the crucible's orifice, thermal conductivity coefficient of the material of
the wheel. The viscosity of the molten alloy also plays a major role.
This chapter is restricted just to the particularities of the melt-spinning
equipment that we used for the preparation of amorphous ribbons with thicknesses
between 20 and 200 IJm
tJm and compositions Fes6
Fe56 CO]
CO? Nb Zr6 Ms
M15 Nb25 8 20 (M = Zr, Ti,
Ta, Mo) and Ndgo - x Fe x A1
Ndgo AI,o- = 20 at. % - 60 at. %, Y == 0 at. % -
10 - y Siy (x =

10 at. % ). The master alloy was crushed into small pieces and introduced in a
%).
quartz crucible in a very close proximity of the Cu rotating wheel. The crucible
exhibits at its lower part a circular or rectangular orifice, depending on the
desired width and thickness of the ribbon. The pieces are re-melted by using
an high-frequency induction coil system. To avoid oxidation, samples were
prepared in an argon atmosphere, after melting in a vacuum (10 - 6 mbar).

9.2.1. 1 Fe-Based alloys

Due to the high viscosity of the molten alloy as well as high melting
temperatures (T mm)) of about 1400 - 1475 'c , and depending on the addition
1400-1475C,
element M, we used quartz crucibles with different rectangular orifices: widths
of O. 3 - 0 . 45 mm and lengths of about 1 mm. The ejection pressure was fixed
to O. 5 - 1 atm depending on the wheel velocity. Ribbons with thicknesses
between 20 and 200 IJm
tJm were obtained by maintaining the orifice size and the
ejection pressure and changing the wheel velocity between 30 and 2.5 m/s.
The widths of the ribbons were between 1. 5 and 3 mm depending on the length
of the crucible rectangular orifice.

9.2. 1. 2 Nd-Fe-Based alloys

Nd-Fe-based ternary alloys have relatively low melting temperatures (about


670C ), so they can be ejected very easily. Due to the high reactivity of Nd,
670'C
the samples were prepared in an Ar atmosphere, after melting in a vacuum
(10- 6 mbar). We used 2 types of orifices: circular with diameters of 0.5 mm for
25 - 80 IJm
tJm ribbons and rectangular (length 1 mm, width O. 5 mm) for 80 - 150 IJm
tJm
ribbons, respectively. The ejection pressure was of about O. 2 - O. 4 atm
(1 atm = 101 ,325 Pa), while the distance between the crucible orifice and the Cu
wheel was of O. 4 mm for circular orifices and O. 8 mm for rectangular ones.
Changing the circumferential wheel velocity from 30 to 2. 5 m/s and keeping
the other parameters constant were prepared ribbons with thicknesses
between 25 and 150 IJm.
tJm.
308 H. Chiriac and N. Lupu

9.2.2 Preparation of Bulk Amorphous Magnetic Rods and Rings by


Casting Techniques
9.2.2.1 Mould Casting Technique
The method, similar to the conventional mould casting method, consists in
applying an overpressure at the upper part of the molten alloy and forces it to
"flow" in a water-cooled Cu mould having different cavities. The master alloy
was melted in a quartz crucible using an induction coil. The crucible holding
system, coil system and the Cu mould are in a chamber connected to vacuum
equipment. During the samples' preparation, an Ar flow is passing
continuously through the chamber to avoid their oxidation. When the pieces of
the master alloy are completely melted, the crucible, with a circular orifice at
the lower part, goes down automatically at about 1 - 2 mm far from the Cu
mould and an overpressure of about 0.5 O. 5 mm is applied
appl ied at the upper part of the
Fes6C07Ni7Zr6Ml sNb 25 B20
molten alloy. Using this equipment, we prepared Fes6C07Ni7Zr6MI
(M=Zr, Ti, Ta, Mo) amorphous cast rods with diameters between 0.5 and
3 mm, Ndgo - x Fe x Al lO
lo (x = 20 - 50) amorphous cast rods with diameters of 1
and 1.5
1. 5 mm and Fes6 C07Nb Ni 7 Zr75
Zru Nb 25 B20 amorphous cast rings with internal
diameter of 6 mm, external diameter of 10 and 1 mm in thickness.
9.2.2.2 Suction Casting Technique
Nd-Fe-based bulk amorphous samples like rods with diameters no larger than
3 mm were obtained by suction casting. The master alloy is arc melted in a
water-cooled Cu crucible and sucked up in a cooled Cu mould. The main
difference between our homemade equipment and suction casting equipments
reported in literature is that both molten chamber and Cu mould chamber are in
the same out chamber, connected to a vacuum pumping system. The pressure
difference required for the suction of the molten alloy in the Cu mould is
created by an overpressure of about O. 3 - O. 5 atm in the out chamber. By
acting as a valve connected also to a vacuum system, the lower part of the Cu
mould is depressurized and the molten alloy is sucked within the mould cavity.
We prepared amorphous cylinders with diameters between 0.6 and 3 mm from
y Siyy alloys (x = 20 - 50; y = 0 - 10), depending on the cavity
Ndgo - x Fe x A1 10 - ySi
profile and alloy composition.

9.2.3 Methods and Techniques Used for Structural and Magnetic


Characterization of Bulk Amorphous Magnetic Alloys
The usual methods to check the amorphous structure of the materials are
X-ray, electron and neutron diffraction. But from diffraction studies, or at least
from X-ray diffraction, which is now the most used method for structural
Bulk Amorphous Magnetic Materials 309

investigation, it is very difficult to distinguish between fully amorphous


structures and crystalline ones in which short-range ordered regions in the
range of 20 A exist. The positions of the atoms in an amorphous solid are not
like in a gas, because the atoms are in contact with each
completely random Iike
other and this imposes a certain degree of structural order. The positions of
the nearest neighbors are generally correlated between them, but in the same
time some randomness could exist. Thus, amorphous sol solids
ids exhibit short-range
structural order but cannot be evidenced by the presence of long-range order.
Moreover, the amorphous structure differs from one composition to the other
Moreover.
as a function of the atoms, type and size as well as the chemical bonding
between atoms. It was proved that two kinds of chemical bonding, metallic
and ionic.
ionic, favor the existence of magnetism in amorphous materials (Coey, eCoey,
1978). The metallic bonding leads to a random dense packed structure with a
coordination number (N) eN) between 8 and 12, while the ionic one favors an
amorphous structure with random coordination structure, and N takes values
between 4 and 8. Despite the analogous values found for the coordination
number in similar crystalline and amorphous materials, the main difference
between their magnetic properties is originated in the non-equivalence of the
atomic positions in amorphous solids,
sol ids, leading to a distribution in magnitude of
the magnetic moments even at 0 K.

9.2.3. 1 X-ray Diffraction


Qualitative information about the structure of Fe-based and Nd-Fe based
ribbons and cast samples were obtained by X-ray diffraction (XRD)
eXRD) using Cu-
Kcx
Ka and Mo-Ka
Mo-Kcx radiations. The structure was checked on both sides as well as
on powders obtained by milling the ribbons, while XRD investigations on cast
samples (rods
erods or rings) were performed using powders obtained after crushing
the rod into small pieces and then milling.
9.2.3.2 Neutron Diffraction
Because of the low penetration depth of X-rays (few efew nm) in Nd-Fe-based
alloys, and due to the limited amount of sample that can be investigated by
X-ray diffraction technique as well as to the controversial amorphous
microstructure supposed to exist in Nd-Fe-based amorphous alloys, we carried
eWannberg et al., 1999). The samples
out neutron diffraction experiments (Wannberg
Ndgo - x Fe x Al lO (x = 20 - 60) ribbons with thicknesses of 25 and
measured were Ndgo-xFexAlloex=20-60)
150 IJm and widths of 2 - 4 mm and Ndso Fe40 Al 10 io rods 1. 5 mm in diameter. The
each, were packed tightly and sealed inside a standard
samples, about 3 - 4 g each.
vanadium container. The sample data were corrected for background and
container scattering, container absorption and sample self-absorption, multiple
and inelastic scattering, and normal ized to an absolute scale. Data were
collected between 15 and 800 K.
The total structure factor, S ( Q), is given by:
SeQ),
310 H. Chiriac and N. Lupu

S(O)=s-c
8CQ)=8-C C9.1)
(9.1)
v-a
where C is the response of the vanadium rod having the same dimensions like
a
the vanadium container, V is the response of the vanadium container, a is the
background and S 8 is the total response of the sample within the vanadium
container. S 0) is invariant in the temperature range, but is strongly
8 (CQ)
dependent on composition and structure. 0 Q== 4nsine/ A represents the wave
vector, e is the diffraction angle, A is the wavelength.
The radial distribution function (RDF)
CRDF) given by
=
00

GCr)=
G(r)= 1 f4n02SCO)sinordO
f4nQ28CQ)sinQrdQ C9.2)
C2n)3
(2n)3 p Or
Qr
o

is the Fourier transform of the total structure factor, S


8 (C0).
Q). Where p is the
number density, r is the interatomic distance.
r) could bring essential information
The calculation and analysis of G (Cr)
about the structure of amorphous materials. Quantitative information about the
inter-atomic distances in the first coordination shell is obtained from the
positions of the maxima of the G Cr). The area del delimited
imited by G CR) gives the
coordination number CN) :
,~

f'.
2
N = f4nr2p(r)dr
4nr pC r)dr C9.3)
'0

r;
where r; is the position of the first maximum of the radial distribution function
and r 0 is the r value for which reduced radial distribution function, G (Cr) r) =
4nr 2 pCr), is zero.
In real life the upper limit of the integral in GCr)
G Cr) is finite, so the Fourier
transform will be truncated. For the SLAD equipment used by us for neutron
diffraction the upper limit is 10.5 A-I. This can be considered a problem
10. 5 A-I.
because for most of the samples, and especially for the amorphous ones, the
structure factor exhibits oscillations at high angles, and in order to apply the
direct Fourier transform on S 8 CO),
CQ ), it should be multiplied by so-called
"modification function". One possible way to avoid these errors is to use an
inverse method.
The reduced radial distribution function, G C r) was obtained from the
Cr)
measured total structure factor S CQ) by using MCGR CMcGreevy and Pusztai,
8 CO)
1988), a program available on SLAD that uses a Monte Carlo method to
"generate" possible G Cr), and then modifies these to fit the data, i. e.
SCO).
8CQ) . :
But the main problem in the analyses of the diffraction data on disordered
systems is the lack of any general method for producing structural models that
agree quantitatively with the data. Also this time, the reverse Monte Carlo
CRMC) method CMcGreevy and Pusztai, 1988) overcomes this problem.
Bulk Amorphous Magnetic Materials 311

In the reverse Monte Carlo simulations reported here, systems containing


2000 - 2500 particles were used. The input experimental data were S ( Q) .
The experimental number density, p, was O. 055 A- 3. Closest-approach
distances were determined on the basis of the corresponding experimental
data. All calculations were started from hard sphere random configurations.
6 6
For reaching equilibrium (the closest fit to the experiment), 6. 5 x
X 10 - 8. 5 x 10

accepted moves had to be completed.


9.2.3.3 Thermal Stability and Glass-Forming Ability
The as-cast amorphous phase is usually metastable, suffering continuous
changes to more stable states. There are 2 kinds of transformations that
strongly affect the physical properties of the amorphous materials: the
structural relaxation and the on set of crystallization. The structural relaxation
appears after applying thermal treatments at temperatures below glass-
transition temperature and is the result of the diminishing of the residual
stresses induced during the preparation process. The diffusion processes
dominate the process and usually physical properties of the amorphous sample
are improved. Crystallization is a very complex phenomenon, which strongly
depends on static and dynamic factors implied in. Different values obtained for
the crystallization temperature (T x) of the amorphous samples prepared by
different methods and having the same composition indicate the formation of
different amorphous structures or just the presence of the impurities. Near T x'
most of the physical properties of the amorphous material, including magnetic
ones, are changing drastically. Thus the value of T x should be known very very
precisely especially for those amorphous materials used in applications.
One of the main factors that affect the glass-forming ability of one metallic
alloy is the thermal stability of the supercooled liquid region (I:i
(11 T x), defined
as the region between glass-transition temperature (T g) and crystallization
temperature (T x), This is dependent to the cooling rate and consequently to
the preparation method.
Previous studies have shown that 2 factors determine the glass-forming
ability: the magnitude of the supercooled liquid region (I:i = T x - T g) and
(11 T x =
the value of the reduced glass-transition temperature ( T g/ T m' where T m is the
melting temperature) .
In subsequent studies, data about glass transition temperature (T 9 ) ,
crystallization and melting temperature, thermal stability of the supercooled
liquid region and the evolution of the crystallization process with the
composition and samples thickness were acquired from calorimetric
measurements (DSC)
COSC) performed at a heating rate of 20 K/s. The difference
between Fe-based and Nd-Fe-based amorphous alloys consists of the position
of the glass-transition temperature relative to the crystallization temperature.
For Fe-based, T g9 is situated below T xx'' and like in most of the previous
studied amorphous alloys, for Nd-Fe-based T 9 cannot be evidenced on DSC
curves, because it is between T x and T m and the large glass-forming ability is
312 H. Chiriac and N. Lupu

crystall ization temperature ( T x I T m)


given by the reduced crystallization m) .

9.2.3.4 Magnetic Measurements

dc-magnetic measurements above room temperature, but not exceeding


1100 K, at applied magnetic fields limited to 1260 kA/m, were carried out
using a homemade vibrating sample magnetometer (VSM). The variation of
the magnetization and coercive field below room temperature (5 - 300 K) was
studied in fields up to 800 kA/m using a superconducting quantum interference
device (SQUID) magnetometer and for maximum applied fields of 7.2 MAim
using an Oxford MagLab VSM. The magnetic field was applied in the axial
direction of the samples. Each sample was thermally demagnetized prior to
recording the magnetic measurements. Zero-field-cooled magnetization
(M ZFC ) was measured on heating in applied field after cooling up to low
temperature in the zero applied field, while the field-cooled magnetization
(M FC ) was measured during warming in the applied field after the field
cooling.
(lJe) and the electrical resistivity (PRT) of the
The magnetic permeability (IJe)
Fe-based amorphous alloys were determined by using an AC fluxmetric method
in a maximum appl
applied Aim,
ied field of 1000 AI m, while the saturation magnetostriction
constant (As) was measured by the small angle magnetisation rotation
( SAMR) method.
AC-susceptibility of Nd-Fe-based amorphous alloys in the temperature
range 4.2 - 300 K was measured by using a very sensitive susceptometer in ac
magnetic fields 160 and 1600 AlmandAim and frequencies ranging from 29. 1 to
291 Hz.

9.2.3.5 Mossbauer Spectroscopy


Magnetic measurements offer information about intrinsic and extrinsic magnetic
properties (coercive field, magnetization, anisotropy, etc.) taking into
consideration the response of either disordered or ordered regions in the
material, so they are strongly related to the structural environments of the
atoms. The only methods so far, that can offer details about the sample's
microstructure as well as its influence on the hyperfine interactions put in
evidence also by magnetic measurements but not proved, are the nuclear
ones: nuclear magnetic resonance (NMR) and M6ssbauer spectroscopy. The
last one is considered a complementary method for diffraction techniques and
allows studying both the atomic scale environments in metals and alloys and
the relaxation or dynamic phenomena that take place in these materials. The
major problem in studying the M6ssbauer spectra of amorphous materials or
generally disordered systems is the fitting procedure that is very complicated
and should take into consideration many intrinsic factors.
M6ssbauer measurements were performed at 300 K on Nd-Fe-based melt-
spun amorphous ribbons in transmission geometry with a constant acceleration
spectrometer, using a 57 Co source in a rhodium matrix. The ribbons, were
Bulk Amorphous Magnetic Materials 313

oriented perpendicularly to the y-beam. In the first approach, the fitting


procedure consisted of using distribution of hyperfine fields linearly correlated
to that of isomer shift, and a quadrupolar component if any.

9.3 Soft Magnetic Bulk Amorphous Alloys

9.3.1 Background
Nowadays, soft magnetic materials represent essential parts in most electric
and electronic equipments and devices governing our life. The most important
requirements for soft magnetic alloys, which are essential for appl ications,
applications,
are: the large values of the saturation magnetization and relative magnetic
permeability, low coercive fields and reduced core losses.
Despite very good soft magnetic properties combined with high strength
and good ductility, conventional amorphous alloys exhibit a major
disadvantage in comparison with crystall ine materials: they cannot be obtained
as bi- or tri-dimensional defined shapes being limited to melt-spun ribbons with
thicknesses up to 50 IJm, wires with diameters no larger than 120 - 150 IJm,
thin films or powders with dimensions of few tens of nm. These limitations are
imposed by the reduced glass-forming ability in Fe-based or Co-based
conventional amorphous alloys, which require high cooling rates of about 105 -
106 K/s
K/ s for amorphization.
For this reason, in the last 5 - 6 years the interest in finding new systems
with large glass-forming ability and good soft magnetic properties has grown.
Large glass-forming ability means high thermal stability of the supercooled
liquid region and consequently the possibility to prepare bulk amorphous
samples with different shapes and dimensions. Equilibrium
Equil ibrium between the large
glass-forming ability and good magnetic properties in the as-cast state should
remain.
There are just a limited number of multicomponent systems in which these
2 requirements, large glass-forming ability and good soft magnetic properties
in the as-cast state, could be satisfied. In the following, we present briefly
these systems, focusing thereafter on the magnetic properties of (Fe,Co,Ni)-
CFe, Co, Ni)-
(Zr,Nb,M)-B
CZr,Nb,M)-B bulk amorphous alloys (M=Ti, CM=Ti, Ta, Mo).
The first results on bulk amorphous alloys with soft magnetic properties
were reported in Fe-(AI,Ga)-(P,C,B,Si)
Fe-CAI,Ga)-Cp,C,B,Si) multicomponent system (Inoue and
FeaD P12 B4 Si 4 basic alloy,
Gook, 1995a). By the dissolution of AI and Ga in a FeaoP12B4Si4
the supercooled liquid region enlarged from 36 to 49 K, making the
preparation of amorphous cylinders with diameters up to 2 mm by mould
casting method possible. Moreover, the magnetic properties indicate a soft
314 H. Chiriac and N. Lupu

magnetic material with saturation magnetization of 1. 1 T, coercive fields


between 6 and 20 Aim, permeabilities of 1.2 1. 2 x 10 4 at 1 kHz and Curie
temperature of 600 K.
The second soft magnetic system reported, in which the preparation of
bulk amorphous alloys is possible, is Fe-TM-B (TM = = transition metal, other
than Fe) multicomponent system (Inoue et al. , 1997a, 1998a; Koshiba et al. ,
1999; Chiriac and Lupu, 1999a). The supercooled liquid region is about 80-
90 K, the largest value reported until now in magnetic amorphous systems.
Bulk amorphous samples like cylinders with diameters up to 5 mm prepared by
mould casting and ribbons with thicknesses up to 200 IJm exhibit good soft
magnetic properties in the as-quenched state: saturation magnetizations of
0.75 - 1.06
1. 06 depending on the nature and amount of TM component, coercive
fields up to 20 Aim, permeabil ities at 1 kHz of 2. 5 x 10 4 and reduced
permeabilities
(A 5 < lOx 10 ). In addition, bulk
6
saturation magnetostriction constants (lis
samples containing Mo or W also exhibit high strength with Vickers hardness
values larger than that ones obtained for hard steels (Inoue et al., 1998b).
Now, the studies in this system are focusing on the explanation of soft
magnetic properties and the interplay between microstructure, glass-forming
ability and magnetic properties.
Recently, the enlargement of the supercooled liquid region in Co-Fe-TM-B
quaternary systems to 45 K by increasing B content to 30 at. % was reported,
but it was not yet possible to prepare bulk amorphous samples by casting
methods (Inoue et al. , 1998c, 1999). The investigations are in progress. It is
worth noting the large values of the permeability of 5500 at 1 MHz obtained for
Co-Fe-Zr-B thick amorphou sribbons, indicating these materials as possible
candidates in the future for high-frequency appl ications.

9.3.2 The Formation of the Amorphous Phase in


(Fe,Co,Ni)7o (Zr,
(Fe,Co,Ni)70 (Zr,Nb,M)
Nb, M) 10B20
IO B2o (M=
(M=Zr,
Zr, Ti, Ta,Mo)
Ta, Mo) Systems

The partial replacement of Fe with a few percent of Co or Ni in Fe-B binary


amorphous alloys increases the saturation magnetization but decreases the
thermal stability of the alloy. The substitution up to 7 at. % of Fe with Ni does
not change the saturation magnetization, but over this value, the decrease of
the magnetization was observed. By the partial replacement of Fe with Co or
Ni no change of the glass-forming ability was observed, while replacing Fe
with Ta, W, Mo, Nb leads to the slight enlargement of the glass-forming
abil ity, the reverse effect is observed for V, Mn, Zr, Pd, AI, Cr (Hagiwara
et ai., 1982). These results suggest that the glass-forming ability is strongly
influenced by physical and chemical factors like: atoms size, electro-
negativity, number of electrons on the outer shell.
Bulk Amorphous Magnetic Materials 315

It is known that the formation of the amorphous phase is possible in


Fe-Zr-B ternary system, the suitable composition being Fe70 ZrlO B20 with a
supercooled Iiquid region 11 D. T x of 83 K (Hagiwara et al., 1982) .
Supplementary, substituting 2 at. % Zr with Nb, 11 D. T x reaches 91 K, the
largest value reported so far for ferromagnetic amorphous systems.
Unfortunately, the softness of this composition is diminished, i. e., large
values for the coercive field and the drastic decrease of the saturation
magnetization (Ma and Inoue, 1999). The large glass-forming ability is
maintained by the substitution of Fe with Co and Ni, the largest value of 68 K
being obtained for the nominal composition Fe S6C0
FeS6 C07Nb
Ni 7Zr
ZrlO10 B20 , which exhibits
in the same time good soft magnetic characteristics: IJo O. 96 T; He =
/.10 M s = 0.96
2A/m; /.1e(l 6
IJe(l kHz)=19,100; A s =10XlO- (lnoueetal.,
C1noueetal., 1997a). WhereIJo Where/.1o
is the magnetic constant, He is caercive field.
What happens with the supercooled liquid region when Co and Ni
substitute Fe and Nb or other refractory metals (Mo, Ti, Ta) substitute Zr?
The response is given by DSC curves recorded at a heating rate of 20 K/s for
Fes6C07NbZr6Ml sNb 25 B20 (M = Zr, Ti, Ta, Mo) amorphous ribbons 25 IJm in
Fe56C07NhZr6M,5NbuB20(M=Zr,
thickness, and shown in Fig. 9.1.9. 1. It is worthwhile to note that DSC curves
exhibit one endothermic peak corresponding to the glass transition. The
existence of one exothermic peak for the amorphous alloys containing
1.5 at. % Zr, Ta or Mo and a shifted peak for Fe S6C0
FeS6 Zr6 Nb u25 Ti,
C07 Ni 7 Zr6 Til 5B
S B20

amorphous alloy indicates one stage crystallization


crystall ization process for the former
alloys and two-stages crystallization
crystall ization for the last one. The substitution of
1.5
1, 5 at. % Zr with Ti, Ta or Mo leads to the increase of the resistance of the
supercooled liquid region against crystallization and consequently the maximum
thickness for which the bulk samples are still amorphous increases about 2
<Table 9. 1). Although the replacement of Zr with Ti, Ta or Mo causes
times (Table

A
}\
T
gl
g1 M=Zr

~
~
I1
~
::i
:::i
I M=Ta
u
<)

' 1
-=~0'"
0)
I /\ M=Ti
x><:
I I
.Y\-
UJ
U.J
M=Mo

IIT, Tx
600 800 1000 1200 1400
Temperature (K)

I
Figure 9. 1 DSC curves of FeS6 Cal
C07 Nh
Nb Zr6 MIS
M 15 Nb25 20 amorphous ribbons (Chiriac and
2 . 5 8 20

Lupu, 1999a).
1999a) .
316 H. Chiriac and N. Lupu

the decrease of the glass transition temperature, the supercooled liquid region
~ T x increases with 10K.
l!.
Table 9.1 Thermal stability and maximum thickness of FeS6 C07Ni7Zr6Nb'5
Fe56 C07Ni7Zr6 Nb,.s M
M'.5
15 8'0
CChiriac and Lupu, 2000al.
amorphous alloys (Chiriac 2000a).

Maximum thickness Thermal stability

I ma , (mml
Cmm) t>.TxCK)
bTx(Kl T G (Kl
TGCK) T x (Kl
T, CK)

M=Zr 1,6
1.6 773 853 80

M=Ta 3 768 853 85

M=Ti 3 733 823 90

M=Mo 3 743 833 90

Another evidence for the formation of the amorphous phase in (Fe, Co,
Ni)70 (Zr,M,Nb),oB
(Zr,M,Nb)lOB2020 alloys is given by X-ray diffraction patterns (Fig. 9.2).
9,2),
They indicate just one broad maximum characteristic to the amorphous phase
and no sharp lines corresponding to crystalline phases for thick ribbons 150 ~m
in thickness, regardless of their composition (Fig.
(Fig, 9.
9, 2a).
2a), The fully amorphous
structure was also found in Fe56 C07NiNb7 Zr7 5Nb2 5B20 bars with diameters up to
1,6 mm and rings with D
1.6 Dint = 6 mm and t = 1 mm as well as in
ext = 10 mm; Din,
De"
bars having diameters up to 3 mm prepared from Fe56 C07Nb Nh Zr6 Nb 25
25 M
M,1 5B20
(M=Ti,Ta,Mo) alloys (Fig. (Fig, 9.2b).
9,2b),

MoKa MoKa
Rod-I.6 Inln
tntn
M=Ta
M~Ta

Ring-/
Ring-I =~ I Inln
tntn
M~Ti
M=Ti
:i :i
~ ~
0 0
'v; 'v;
c c
<l) <l)

E M~Mo
M=Mo E
~
Ribbon- 85 l1 tn

M=Zr
M~Zr Ribbon-30 ~In
I1tn

15 20 25 30 35 40 15 20 25 30 35 40
2en
2en 2en 2en
(a) (b)

Figure 9. 2 XRD patterns of FeS6


Fe56 C07Nh Zr6 MIS
M 15 Nb 25 8'0 ribbons 150 j.Jm
jJm in thickness (al
Ca)
and Fe56C07NhZr7 5Nb, s8,o
FeS6C07NbZr7 sNb, 58'0 ribbons, cast rods and rings (bl,Cb).
Bulk Amorphous Magnetic Materials 317

The existence of a single amorphous phase and no crystalline traces in the


Fe56 C0 5 Nb2 5 8 20 bulk amorphous samples is evidenced also by
Co?7 Nb Zr6 M 1 5Nb
thermomagnetic measurements, as it can be seen in Fig. 9. 3 for the basic
= Zr. The first decrease of the magnetization to zero with the
alloy with M =
temperature increase (solid symbols curve) gives the value of the Curie
temperature of the amorphous phase (about 554 K) and proves the
amorphicity of the sample. The magnetization starts to increase by the on set
of crystallization (about 860 K), and reaches a maximum and decreases again
to zero when the fully crystalline
crystall ine sample becomes paramagnetic, indicating a
value around 1100 K for the Curie temperature of the crystalline
crystall ine alloy. The
second graph (open symbols curve) is characteristic for the crystallized alloy.
Thus, the magnetization increases when the sample suffers a phase transition
from amorphous to crystalline.
crystall ine.
1.2
0-0- H 200 kAhn
ext= 1200
Hext=1
1.0

0.8
E
1
i 0.6
0.6

0.4

0.2

0
300 400 500 600 700 800 900 1000 1100
Temperature (K)

Figure 9. 3 Thermomagnetic curves of Fes6Co7NhZruNb25B2o


FeS6 Co, Nb Zru Nb25 8'0 amorphous (solid
Csolid
symbols) and crystalline (open
Copen symbols) alloys, respectively (Chiriac
CChiriac and Lupu, 2001a).

9.3.3 Electric Resistivity of Fe-Based Bulk Amorphous Alloys

Electric properties of amorphous materials are usually very sensitive to


structural disorder, stress relaxation, differences between the free volumes as
a function of the cooling rate or composition. Consequently, the study of the
electrical characteristics in amorphous materials is very difficult, and usually
just relative measurements can be done. From the previous results we can
notice that the electrical resistivity of the amorphous materials ranges between
1 x 108 and 3 x 1O- 8 Qm.
In Fig. 9.4 the temperature variation of the electric resistance for the
318 H. Chiriac and N. Lupu

amorphous ribbon 25 IJm ~m in thickness with nominal composition


Fe56 Co,
Fes6 C07Nb Zru Nbu B20 is presented.
10.25
10.20
10.15
.-----
_
.--"--- ..../ \ /
............-"
.-.......... ~
...."

'%
10.10 ,r
iP""
a 10.05
~ 10.00
Ci2"
/
/ ""
<1'"/__
,// Amorphous
-o-Crystalline
9.95
/0
0/0
9.90
9.85 '::3~00;:---4-:-:0~0----'5:-':0-;:-0---:6~00;:---7=:0~0---:8:-':0-;:-0
300 400 500 600 700 800 --::-:'900
900
Temperature (K)

Figure 9.4 Temperature dependence of the electrical resistance for


Fes6Co,Ni,Zr, sNb25 B
8 20 amorphous and crystalline alloys.

The resistance increases linearly with a temperature below the


crystallization temperature (T x = 853 K), and then starts to decrease to the
value corresponding to the crystalline alloy. At high temperatures the curves
corresponding to amorphous phase and crystalline
crystall ine material, respectively, are
superposing. It is worthwhile to note the essential difference at room
temperature between the electrical resistance of the amorphous and crystalline
materials with the same composition. This difference is diminished by
increasing the temperature, due to the relaxation processes and the on set of
crystall ization in amorphous material. Thus, the temperature variation of the
crystallization
electrical resistance gives useful and accurate information about the relaxation
and crystallization processes in amorphous materials.
We calculated the room temperature electric resistivity C
(PRT)
PRT) of
CFe,Co,Nj),oCZr,M,Nb)10B20
(Fe, Co, Ni)70 (Zr, M, Nb) 10 B20 amorphous alloys from the measured electrical
resistance (Table 9.2). One can see that PRT increases by the addition of Ti,
Ta or Mo as a result of the increase of the structural disorder due to the
different dimensions of the atoms.

Table 9. 2 Room temperature electrical resistivity of FeS6 Co, Ni, Zr6 M


M1.5
15 Nb 25 B
8 20 amorphous
alloys.

M=Zr 151
1. 51
M=Ti 2.08
M=Ta 1.87
1. 87
M=Mo 176
1. 76
Bulk Amorphous Magnetic Materials 319

9.3.4 (Fe,Co,Ni)7o(Zr,Nb,M) JO B2o Bulk


Magnetic Properties of CFe,Co,Ni)7oCZr,Nb,M)IOB2o
Amorphous Alloys in the As-quenched State
9.3.4.1 Composition Influence
The change of the composition of the Fe56 C0 Co?7 Nb Zrn
Zr?5 Nbu25 B20 basic alloy
causes not just the increase of the glass-forming ability shown in the previous
section, but also the modification of the soft magnetic properties as it can be
seen in Table 9. 3. While the saturation magnetization does not change
significantly by introduction of Ti, Ta or Mo instead of Zr in the basic alloy, the
magnetic permeability decreases from 21,500 for the Fe56C07NbZrnNb25B20
Fe56Co?NbZruNbuB20
amorphous ribbons having 35 IJm in thickness to 17, 000 for the
Fe56 C0
Co?7Nb Zr6 Nb25
u M0 15 B20 amorphous ribbons with the same thickness. The
Curie temperature increases and the coercive field decreases by the
substitution of Zr with Ti, Ta or Mo.
Table 9.3 Soft magnetic characteristics of the Fe56 Co? Nb Zr6 Nb25
FeS6 COl M1.5 8 20 (M
2 . 5 MIS CM = Zr,
Zr. Ti, Ta,
Mo) amorphous ribbons having 35IJm in thickness (Chiriac
CChiriac and Lupu,Lupu. 2000a).
~o
fJo M s (T) He (Aim)
HcCA/m) ~e (500Hz)
fJ.C500Hz)

M=Zr 1.01 95 21,500


21.500 554

M=Ta 0.89 6.06


606 18,000
18.000 560

M=Ti 1..06
06 66.11 19,000 603

M=Mo .07
1.07 743
7.43 17,000 560

This behavior could also be explained on the basis of the increase of the
thermal stability of the supercooled liquid region. The crystallization energy,
determined by the area under the exothermic peak on DSC curves, is larger
for amorphous alloys with Ti and Mo, suggesting a higher degree of structural
disorder and consequently the development of different magnetic structures is
more stable.
Table 9. 4 presents the frequency dependence of the magnetic
permeability for Fe56 C07Nb Zr6 M155NbuB20
Fe56Co?NbZr6Ml Nb25 B20 amorphous alloys.
Table 9.4 Magnetic permeability as a funciton of frequency for Fe56 Co? Nb Zr6 MIS
FeS6 COl M1.5 Nb25 8 20
(M
CM== Zr,
Zr. Ti,
Ti. Ta, Mo) amorphous alloys.
Ta.Mo)amorphous
v=5 Hz v= 10 Hz v =50 Hz
v=50 v =500 Hz v= 1 kHz v = 10 kHz
v=
M=Zr 33,000
33.000 31,600
31.600 30,900
30.900 21,500
21.500 14,900
14.900 5,400

M=Ta 21,300
21.300 21,600 19,500
19.500 18,000 16,000
16.000 11,000
11.000

M=Ti 20,400
20.400 20,900
20.900 19,950
19.950 19,000 18,000
18.000 8,000

M=Mo 20,000
20.000 21,000 20,700
20.700 17,000 15,750
15.750 7,300
7.300
320 H. Chiriac and N. Lupu

The magnetic permeability of the basic Fes6Co?Ni?Zr6Mj


Fe56 COl Nil Zr6 MISsNb
Nb 2 sB
5B20
amorphous alloy increases by about 50 % by decreasing the frequency from
500 to 5 Hz and decreases about 4 times with the frequency increase to
10 kHz. By the partial substitution of Zr with other additional elements, the
10kHz.
variation of the magnetic permeability as a function of applied frequency is less
pronounced, indicating these bulk amorphous alloys with additions as suitable
applications.
candidates for appl ications.
The large values of the magnetic permeability obtained for
FeS6 Co? Nb
Fe56 COl Nil Zr6 M 15 Nb u B20 amorphous alloys are closely related to the high
1 5Nb
electrical resistivities (PRT) and to the relatively small saturation
magnetostriction constant (i\ CA s) measured in the as-cast state. i\As ranges from
7 x 10- 6 for Fe56
FeS6 COl Zru Nb uu B20 amorphous alloys to 10 x 10- 6 for the
Co? Nb Zr75
amorphous alloys having Ti, Ta or Mo in composition. The small values of i\Ass
explain also the reduced values of the coercive field.

9.3.4.2 Thickness Dependence

The improvement of the magnetic properties of the Fe56 COlCo? Nb


Nil Zr6 M Nb 2 5B
M,155Nb S B20

amorphous alloys is obtained by choosing the suitable composition or by


increasing the thickness of the samples, namely by attaining a high degree of
structural relaxation in the as-cast state. Figure 9.5 shows the behavior of the
magnetic permeabilities at 500 Hz as a function of thickness for
FeS6 Co?
Fe56 Nb Zr75
COl Nil Zru Nbu B20 melt-spun ribbons, and cast rods. The increase of the
amorphous ribbon' s thickness leads to the structural relaxation and
consequently the permeability increases. The thicker amorphous ribbons, the
higher degree of structural relaxation and the better magnetic properties. The
role of the eddy current in the amorphous ribbons with thicknesses up to 150
3.0

f
o Rod d=0.5 mm 251.lIn
Ribbon 25 J.lm xl
x I mm
o Rodd=1.6mm
Rod d=1.6mm Ribbon601.lInxlmm
Ribbon 60J.lmX I mm
d=3 mm~.
2.5
o
b. Rod d"3mm .... Ribb" 165,,,x
Ribbon 165 f.lm xlI mm

2.0
" /i
.,.
o
x 1.5
~

::t
'1\ \'
\ \ 0

I \ .\
.I "\f.~:~\
~."".:~:
"'\~'.
1.0

5.0
II "
" . . . .'. .
A___ r
.r /'l
~.;
.'... '~_
-~ :-.
.:> .....~ ~!_..
"............. " ;)o ............~!_ .
'-._.-r-'..l.-o./.
"-"-,,----"-+"-... ~==::~ 0' ".:.'no
"'. o~_~
'- ~ll'l:l
o bo:P'
I 10 100
Applied field (Aim)

Figure 9. 5 Cat 500 Hz) of FeS6C07NbZr7


Magnetic permeabilities (at FeS6Co, Ni,Zr, sNb,s8,o
5 Nb zs 8 20 melt-
me11-
spun ribbons and cast rods.
Bulk Amorphous Magnetic Materials 321

IJm is very small and hence it could be neglected. The crystall ization on set in
ribbons with thicknesses larger than 150 IJm results in the decrease of the
magnetic permeability and in the displacement of the maximum of the
permeability towards high-applied fields. The magnetic permeability of the
FeS6 COl
Fe56 Co? Nb Zrl
Zr? 5Nb
5 Nb 25
u B 20 amorphous cast rods with diameters up to 1. 6 mm is
8 20
smaller than the values obtained for melt-spun amorphous ribbons. This
behavior is related to the development of different magnetic domain structures
in rods in comparison with ribbons due to the different cooling rates used in the
preparation process and the difference between the amorphous samples
dimensions. For Fe56ColNilZruNb2
FeS6 Co? Nb Zru Nb 2 5820
5B20 cast rod 3 mm in diameter, which is
crystalline in the as-cast state, the magnetic permeability is very small. Thus,
the good soft magnetic properties of Fe56 FeS6 CO] M,1 5Nb
CO? Nb Zr6 M 5 Nb2
2 S
5B 20 bulk amorphous
8 20

alloys are strongly related to the existence of the amorphous phase.


It is important to notice that at 10 kHz the magnetic permeability is the
same for Fe56
FeS6 COl
Co? Ni?
Nb Zr6 M 15 Nb u 8 20
25 B 20 amorphous alloys in the shape of ribbons
and bulk samples (rods or rings), while at 500 Hz they differ with about 35 % .
This observation suggests bulk amorphous alloys as good for application at
intermediary frequencies.

9.3.5 Magnetic Properties of Fe-Zr-B-Based Bulk Amorphous


Alloys Versus Annealing
Structural relaxation of FeS6
Fe56 Co?
COl Nb Zr6 M 15 Nbu25 B
1 5Nb 8 20
20 amorphous alloys can also

be reached by annealing the amorphous samples at temperatures below glass


9. 6 presents the evolution of the magnetic
transition temperature ( T g), Figure 9.6
permeability and magnetization of FeS6 Co? Nb
Fe56 COl Nil Zru Nbu25 B8 20
20 amorphous thin

ribbons as a function of the annealing


anneal ing temperature. Treatments applied
appl ied below
753 K (20 K below T gg)) leave the sample in the amorphous state and lead to
the increase of the magnetic permeability while the magnetization remains
almost the same as in the as-cast state. The subsequent increase of the
annealing temperature results in the on set of the nucleation leading to the
decrease of the magnetic permeability and the increase of the magnetization.
At the same time, by increasing the annealing temperature the magnetization
decreases but does not become zero, indicating a mixture of amorphous and
crystall ine phases.
crystalline
The coercive field initially exhibits a decrease with temperature and starts
to increase when annealing temperature exceeds T x' this increase being
pronounced for the basic alloy (M = = Zr), as it can be seen in Fig. 9. 7. The
precipitation of crystalline phases results in a sudden increase of the coercive
field (He)'
We observed that thermal annealing applied to the thick amorphou
322 H. Chiriac and N. Lupu

3.0
- as-cast
2.5
" 553 K
753 K
.,. 2.0
~
".
0

~
xx 1.5
763 K
823 K
0 938 K
938K
:t
::t
1.0
.-
/+
/+ //
5.0
o-
I
/ 1000

1.2

1.0

0.8

E
~ 0.6
0.6
~ 0.4

0.2 763K
753 K
oL..-_-----L_
0 _-'-----~~~~~
300 400 500 600 700
Temperature (K)

Figure 9.6 Magnetic permeability and thermomagnetic curves vs. annealing


temperature for FeS6
Fe56 Co)
Co, Ni)
Ni, Zr75 Nb
Nb,2 S8
58'0
20 amorphous alloys (Chiriac and Lupu, 2001 a) .

80

M=Zr
60 M=Ti
...
A M=Ta
,.
... M=Mo

~ 40
$
~
20

o 300 400 500 600 700 800 900 1000


Annealing temperature (K)

Figure 9.
9.77 Coercive field vs. annealing temperature for Fe56
FeS6 Co)
Co, Nil
Ni, Zr6 Nb
Nb,2 SM,
5M, S8
58'0
20

\.1m in thickness (Chiriac and Lupu, 2000a).


amorphous ribbons having 35 iJm
Bulk Amorphous Magnetic Materials 323

sribbons and cast amorphous rods does not change the magnetic properties
due to the more relaxed structure attained in the as-cast state.
In the following, let's see how the thermal treatment combined with an external
magnetic field affect the magnetic behavior of Fes6 CO]COy Nb Zr6 Ms Nb25 2 . s 8 20 bulk
amorphous alloys. Figure 9. 8 shows the modification of the shape of the
hysteresis loops after thermal and thermomagnetic treatments for one thick
amorphous ribbon (120 IJm) prepared from the basic alloy Fes6 CO]
COy Nb Zru
Zr75 Nbzs 8 20 .
2 5820
The direction of the applied magnetic field is relative to the long axis of the
ribbon.
1.00

0.75
-as-cast
0.75 ----300"/748K
----300"/748 K
withoutH
withoutHext
exl
/
_ -1- -::: :.. :-:-;;,:!!,;;,r.-..=...
.-
:.~
;::-
=
0.50 ........300"/748 K ' I
withoutHex,l! I
-._.- 300"/748 K I
0.25 withoutHex,l-1
F
~ 0o
_ _ _ --L

I I.:
~ -{).25
1-0.25 I.
I.:
i ..A
1/:
i
/ I ;/
--{).50
-0.50 / ,-r
/' /..... . I
-{).75 .~- - ....: ..~ ..~..::.. ~.. ~..::- -
-1.00 '-- -'-
--'- -'-
---I.. -'---
.1....- --'
---.J

-20 -10 10 20

Figure 9.8 The hysteresis loop shape dependence on the treatment for
FeS6 Co,
Fe56 Co? Nh
Ni, Zru
Zr7.5 Nb2.5
Nb25 8 20 amorphous alloys.

One observes that the rectangularity of the hysteresis loop as well as the
saturation magnetization and coercive field are changing with the direction of
the applied magnetic field. While the thermal treatment without a magnetic
= 773 K) leads to the slight increase of the coercive field and
field below T g9 ( =773
to the elongation of the hysteresis loop. The presence of an external magnetic
field results in better magnetic properties, i. e.,e. , the better softness of the
amorphous material. The best response is obtained when the applied field is
parallel with the long axis of the ribbon because of the easy axis induced on the
parallel direction with the applied field. When the external field is
perpendicular to the ribbon long axis the induced anisotropy is smaller, owing
to the competition of the magnetoelastic anisotropy induced during the melt-
spinning preparation process and oriented parallel with the long axis.
Consequently, the coercive field increases. Further experimental work along
this line is expected in the near future.
324 H. Chiriac and N. Lupu

9. 4 High-Coercivity Bulk Amorphous Alloys

9.4.1 Background
The ferromagnetism of amorphous alloys was widely investigated in the last 40
years, but in spite of that a clear dimension on the complexity of the magnetic
structures developed in these materials does not exist.
The magnetism of rare earth (RE) - Fe glassy alloys produced by rapid
quenching from the melt, has been a great stimulus for research in their
preparation and behavior. These alloys form a variety of magnetic structures
ranging from ferromagnetic to speromagnetic and sperimagnetic, depending on
the relative magnitude of random magnetic anisotropy (RMA) with respect to
the exchange energy (Coey, 1978; Moorjani and Coey, 1984).
The high coercivity, over 10 kOe at room temperature, observed by
Croat in melt-spun rare earth - Fe (RE = Nd or Pr) amorphous alloys is perhaps
(RE=Nd
the first evidence for the existence of an unknown metastable hard magnetic
phase in the light rare earth-transition metal binary amorphous systems
(Croat, 1982). The large values of the coercivity obtained for these alloys can
be ascribed to the existence of a new metastable glassy or nanocrystalline
phase, so called AI Al phase, which exhibits a high anisotropy field (Givord
et al. , 1992). The magnetic properties of RE-Fe binary amorphous alloys are
also of technological interest because of their wide range of applications as
permanent magnets, magnetostrictive materials, and magneto-optical
recording media.
Recently, it has been reported that Nd-TM-AI and Nd-Fe-Si ternary
amorphous alloys are formed in a wide range of compositions by melt spinning
and mould casting and exhibit large coercivities at room temperature (He
et aI.,
al. , 1994; Inoue et aI.,
al., 1997b; Chiriac and Lupu, 1999b; Wang et aI., al. ,
aI., 1999; 0' Connor et al. , 2000). Their magnetic behavior
1999; Fan et al.,
indicates that they are structurally glasses but magnetically granular with
coercive fields as high as 8 T at low temperatures and very large applied
appl ied fields
Ortega-Hertogs,, 2000). Coercive field is believed
( Ortega-Hertogs bel ieved to result from the
development of a mixture of different local atomic structures (Matsubara
et al., 1997). These results are in contradiction to those found in NdFeB
conventional ternary amorphous alloys, in which the amorphous microstructure
gives rise to soft magnetic characteristics with negligible
negl igible coercive fields
(Mishra, 1986), but they are in agreement with the results obtained on
amorphous Nd-Fe binary alloys (Croat, 1981). It is worthwhile to notice that
the addition of AI to Nd-Fe-B amorphous alloys improves the coercivity by
Bulk Amorphous Magnetic Materials 325

modifying the microstructure (Knoch et al. ai.,, 1989).


In the following, we present the magnetic behavior of Nd90
go -- xx Fe x (AI,
(AI,Si)lo
Si) 10
(x = 20 at. % - 60 at. %) bulk amorphous alloys trying to explain this behavior
considering the structural and magnetic experiments evidences and simulated
topological structures in the hypothesis of a random atomic configuration. Ndso
Fe40 AI lO will be considered in the following the basic composition, from which
Al IO
the others are derived.

9.4.2 Glass-Forming Ability in Nd-Fe-(AI,SD


Nd-Fe-(Al,Si) Systems

Recent investigations show that the glass forming ability increases by the
addition of elements like AI, Si, Ge to Nd-Fe binary systems. Consequently,
the formation of the amorphous phase is extended to a wide range of
compositions. For a Nd-Fe-AI ternary system the formation of the amorphous
phase is possible in the composition ranges: 0 at. % - 90 at. % Fe and
Oat. %-93 at. % AI (Fig. 9.9).
Oat.%-93at.%
AI 0 Amorphous
() Amorphous+crystalline
~ Crystalline
~
-b
00
Fe (at.%) 'h AI (at.%)

Fe 10 20 30 40 50 60 70 80 90 Nd
Nd (at.%)

Figure 9.9 The composition range in which the amorphous phase is formed
in Nd-Fe-AI ternary system (Inoue et al..
al. , 1997b).
, 997b) .

This relatively wide range of compositions is not a surprise, because the


corresponding composition ranges in binary Nd-Fe and Nd-AI systems. The
increase of the glass-forming ability,
ability. which makes the preparation of bulk
amorphous samples in Nd-Fe-AI ternary system possible, is due to the
dissolution of AI in Nd-Fe binary system.
The large glass-forming ability of Ndso Fe40 (AI, Si) 10
10 alloys is due to the
high values of the reduced crystallization temperature (T x/x / T m) (Table 9. 5) ,
because DSC curves indicate only one exothermic peak due to crystallization
and neither endothermic reaction due to the glass transition nor the
supercooled liquid region in the temperature range before crystallization
(Chiriac et al. , 2001 b). One observes that the total replacement of AI with Si
326 H. Chiriac and N. Lupu

results in the decrease of the glass-forming ability and consequently the ability
to fabricate bulk cast samples with fully amorphous structure decreases, as it
can be seen in Table 9.5 9. 5 and Fig. 9.10.
9. 10. DSC curves show no exothermic
peak of crystallization for Ndso Fe40 SilO melt spun ribbons thicker than 1 00 ~m
100
and cast rods, indicating fully crystallized
crystall ized samples. It is important to note that
the largest values of the reduced crystallization temperature are obtained for
NdsoFe4oAl1O amorphous ribbon 105 ~'m
NdsoFe4oAIlO ~m thickness and respectively NdsoFe4oSilO
amorphous ribbon 50 ~m thickness, which release also the largest amounts of
heat during crystall
crystallization
ization suggesting the highest degree of structural disorder
in these two amorphous ribbons.
Table 9.5 Thermal stability and X-ray diffraction results for Nd 50 CY = 0, 10)
so Fe,o Al lO - y Siy (y
CLupu and Chiriac, 2001 a) .
melt spun ribbons (Lupu

Alloy's Thermal stability


Thickness(lJm
ThicknessC IJm)) - Structure
composition T x (K)
CK) T m (K)
CK) Tx/T m
30 812 931 0.87
087 amorphous

50 804 928 087 amorphous

so Fe,o Al lO
Nd50 85 806 927 0.87 amorphous
105 818 929 0.88
088 amorphous
142 803 926 087 amorphous

30 737 946 0.78


078 amorphous

50 780 943 0.83


083 amorphous

so Fe,o Si
Nd50 Silo
10 85 753 946 0.80
080 amorphous
105 947 crystalline
142 947 crystalline

Ton I
Ton)

=
~

~
:J

' ~--------,
(l)
OJ
.J::
-'=
o
~ y=O
t.:3
w

soo
500 600 700 800 900 1000
Temperature (K)

Figure 9. 10 DSC curves of NdNdso CY = 0, 5, 10) cast rods 1.5 mm in


50 Fe,o Al lO - ySiy (y

diameter (Chiriac
CChiriac and Lupu, 2001 a) .
2001a).
Bulk Amorphous Magnetic Materials 327

The increase of the Nd content, keeping AI to 10 at. at, %, results in the


enlargement of the glass-forming ability,
ability. The largest diameter obtained for an
amorphous rod is 3 mm for nominal composition Nd55 Fe35 Al lO lO ,, and glass
forming ability decreases by increasing the Fe content. These alloys
crystallize through a single exothermic reaction and exhibit very large values
for the reduced crystallization temperature (T CT xx / T m) of 0.87 to 0.92,
0,92, as it can
be seen in Fig, 9. 11. The crystallcrystallization
ization energy (the Cthe area under the
exothermic peak) is higher for the melt-spun ribbons containing a higher
amount of Fe in comparison with the rods due to the more disordered
amorphous structure developed in the first ones, ones. The crystallization energy
doesn't change significantly as a function of the Fe content for Nd go - x Fe x Al lO
Ndgo-xFexAllO
thin ribbons 30 \.1m
IJm thickness, whereas for cast rods it strongly decreases with
the increase of the Fe content as a consequence of the decrease of the glass-
forming ability,
ability. This behavior is explained by the insufficient cooling rates
assured by the casting techniques for the alloys with larger contents of Fe and
indicates a strong correlation between the composition, the cooling cool ing rate and
the glass-forming ability for Nd-Fe-based ternary amorphous alloys, alloys.
Melt spun ribbons-25 11m
~m Cast rods Tml Tm
x=35 x=35 1.5mm I I

S'
~
:::> x=40 S' x=40 1.5 mm
~ ~
u
<,)
,2
.~
'
'
....
"
..t:
x=45 E

"0
..t:
.<::
x=45 1.5 mm
,/" -- l
nl
I
~0
x>< x>< IZ:'I I TX2 Tm , \
LJ..l
U..l x=50 LJ..l
U..l x=50 I mm ./'
I, r,z:, IT,
IT, \f
600 700 800 900 600 700 800 900
Temperature (K) Temperature (K)
(a) (b)

Figure 9. 11 DSC curves plotted for Ndgo x Fe, Al 10


go -- , lo melt-spun ribbons and cast rods as a
function of Fe content (Lupu et al. , 2001 b) ,

9.4.3 X-ray and Neutron Diffraction Experiments on


Nd90-xFexAllO-ySiy Bulk Metallic Alloys
Nd90-xFexAIIO-ySiy

Neutron and X-ray scattering experiments have been extensively employed to


elucidate the structure of amorphous materials, even if they usually yield a
one-dimensional description of the three-dimensional distribution of the atoms
in the amorphous alloys.
alloys, What kind of information do these two techniques
offer us in the case of complicated Nd-Fe-CAI,Sj)
Nd-Fe-CAI, Si) bulk amorphous alloys?
328 H. Chiriac and N. Lupu

9.4.3. 1 X-ray Analysis

X-ray diffraction patterns show no Bragg lines due to crystalline phases for
ribbons with thicknesses between 30 and 142 ~m and nominal composition Nd50
so
Fe 4o AI 1o (Fig. 9.12), as well as for rods up to 2 mm in diameter.
Fe4oAI,o(Fig.

Cu Ka 142 !lID
14211m

20 25 30 35 40 45 50 55 60
2en
28 (')

Figure 9. 12 X-ray diffraction patterns for Ndso Fe,o Al lo melt-spun ribbons.

XRD patterns also indicate fully amorphous structures for Ndgo-


Ndgo - x Fe x AI,o
Al lO (( x =
20 at. % -50 at. %) ribbons with thicknesses below 100 ~m (Fig. 9.13) and
for rods with diameters less than 1 mm, regardless of their Fe content. The
increase of the Fe content over 45 at. % results in the on set of crystallization
in melt-spun ribbons thicker than 100 ~m, while for 60 at. % Fe the melt-spun
ribbons 30 ~m in the as-cast state are crystalline.
crystall ine.

20 30 40 50 60 70 80
2e (0)
28 (')

Figure 9. 13 X-ray diffraction patterns, as a function of x, for Ndgo - x Fe x Al lOlo melt-


spun ribbons.

The X-ray diagrams indicate a fully amorphous structure for Nd50


so Fe40 Al lO - ySiy
melt-spun ribbons with thicknesses less than 100 100 ~m, in all ranges of
compositions, and also for bars with diameters less than 1. 8 mm prepared
Bulk Amorphous Magnetic Materials 329

from alloys, which contain up to 2 at % Si. The ribbons thicker than 100 IJm,
at%
lo - y Siy bars (y ~
the Ndso Fe40 Al lO :S;;; 2 at%) with diameters between 1. 6 and
3 mm and the bars with diameters up to 1. 6 mm prepared from alloys
containing 2 at. % - 5 at. % Si exhibit weak Bragg lines of ex-Nd and Nd2 Fe17 Fel7
superposed to the halo-pattern, typical for the amorphous phase (Fig. 9. 14) .
Moreover, the Ndso Fe40 SilO Silo ribbons having more than 100 IJm and bars with
diameters less than 1.6 mm present also very weak Bragg Iines due to ex-Fe.
The Ndso Fe40 A1 IO - y Si
AI 1o Siyy bars containing more than 5 at. % Si have a fully
crystall ine structure.

Cu Ka a-Nd
* 2: 17 phase
a-Fe

;:;
~
0
';j;
<:
~ *
<:
1051lm

90 Ilm
70 Ilm
251lm

20 30 40 50 60 70 80
2en
Figure 9. 14 X-ray diffraction patterns of the Nd50
so Fe,o SilO melt-spun ribbons having
Fe40 Si,o
different thicknesses (Chiriac et al. , 2001 c) .

9.4.3.2 Neutron Diffraction


The investigation of the disordered materials by neutron diffraction has few
advantages in comparison with the X-ray diffraction: the neutrons scattering
length is almost independent on the Q Q ( = 4n
41T sine/II)
sine / i\) value, and thus it is
easier to determine the total structure factor S ( Q) for large Q values with
accuracy; the neutron scattering length is independent of the atomic number of
the elements, so light atoms like AI, Si or 0 can be identified by neutron
diffraction but no so easy by X-ray diffraction.
In Fig. 9. 15 the neutron diffraction patterns taken at room temperature for
Ndgo - x Fe lo (x = 35, 40, 50) melt-spun ribbons with thicknesses of 25 and
x Al lO
FexAl
120 IJm are presented, respectively. Due to the decrease of the glass-forming
Nd4oFesoAlIO ribbon 120 IJm, cannot be obtained fully amorphous.
ability, the Nd4oFesoAI,o
The split first peak and the pronounced structure at higher angles are features
not present in the patterns for the other amorphous alloys with dense random
packing disordered structures but they were observed in Nd-Fe binary
330 H. Chiriac and N. Lupu

3.5
3.S 1.2
Nd 5S FeJ5 AI IO
Nd5SFe35Alio
3.0
~2S/lm
-25/.un 0.8 -25111n
-2S/l1n
2.5
2.S 120111n
120/lrn 120 111n
120/l1n
0.4
6i
&2.0
2.0
C;)
C;;)
o
1.5
I.S

1.0 -0.4

5"'------o----'-----l~--',----,L----l
O. S ----::---'----':----',---,-L----' -0.8 2
o 2 4 6 8 10 -0.8 33 44 5
S 66 7
7 88 99 10
10 I/
II
Q(I/A) r (A)
(a) (b)

3.S
3.5 2.0
NdSoFe40Alio
3.0 1.5
I.S
- 25 111n
-2S/lrn - 25 111n
-2S/lrn
2.5
2.S 120
120/l1n
11 111 1.0 120111n
120/lrn

6i
&2.0
C;)
2.0
C;;)
i O.S

1.5
I.S o
1.0 -0.S ,
0.5'----_--'---_----'--_ _'----_--'---_-'----'
O.S"--_--'---_----'--_-----.J"--_--',-_-'---'
o 4 6 8 10 -1.0 2 3 4 5
S 6 7 8 9 10 I11J
Q(I/A) r (A)
(c) (d)
5.5 1.0
5.0 Nd 40 FesoAl
FeSOAI\o
io 0.8
4.5
~
- 25111n
2Sl!rn 0.6 - 25111n
2S/lrn
4.0
3.5
120111n
120/l1n 0.4
~ 3.0 S 0.2
C)j
C;;) 2.5 c.J
c.:J 0
2.0 -0.2
1.5
-0.4
1.0
0.5 -0.6

o 2 4 6 88 10 -0.8 2 6 -7=---;8:--~9:--.,';1
!;---:3:----o4--=-S--7
=--=3:----;4--=-5--0- -7=---;8:----:!9~.,.,10,------711
0,.----,'11
Q(I/A) r (A)
(e) (f)

Figure 9. 15 Neutron diffraction pattern as a function of wavevector, Q Q = 41Tsine/ ii,


i\,
and reduced radial distribution function G ( r) = 41Tr 2 p ( r) for Nd90 - , Fe, Al lo melt-spun
ribbons (Lupu et al. , 2002).

amorphous alloys (Alperin et al. , 1979). The split


spl it is more pronounced for the
thicker ribbon than for the thin ones. We suppose this behavior is related to
the more relaxed structure developed in the thick ribbons. The reduced radial
), obtained by Fourier transform of the total
distribution functions, G ( rr),
structure factor S ( Q), show first neighbors peaks at 2. 54, 2. 85, 3. 36 and
4. 20 A. The peak at 2. 54 A is at the position expected for the nearest
Bulk Amorphous Magnetic Materials 331

neighbor Fe atoms in the dense random packing (DRP) model; the other
peaks cannot be correlated to combinations of the radii of Nd (1. 82 A),
AI (1.43 A) or Fe (1. 27 A) atoms as predicted by the DRP model. This
disagreement, which is particular to the case of the present Iight
light rare-earth
RE-TM alloys, could be ascribed to the development of a new type of
disordered structure. Thus, one suggests that the structure consists in a dense
random packing of nanometer-sized atomic clusters, the size being dependent
on the thickness of the ribbon, i. e. the melt-spinning conditions.
In agreement with the XRD results, the increase of the Fe content results
in the precipitation of the crystalline phases, as it can be observed in
Fig. 9. 16 for Nd
9.16 30 Fe60 Al lO melt-spun thin ribbons.
Nd30FesoAlIO

",,,. ""'"......
""" x=60
4
--x=50
...... _. x=40
---- .....
-.------ x=35
;
Ii:1
1; ... --_..
.. x=30
33 ::: i!
.. x=20
]
i
!!
!
Sl I
C5"
2
: i~
Of:
I
!:

o 2 3 4 5 6 7 8 9 10
Q (IIA)
(I/A)

Figure 9.16 SCO)


9. 16 Total structure factor S CQ) as a function of x, for Nd90
oo -- xx Fe xx AI\o
Al lO melt-
spun ribbons.

The interatomic distances and the coordination numbers, as a function of


x, were calculated from the reduced radial distribution function G (r), and are
given in Table 9.6.
Table 9.6 The interatomic distances Cr) Cr) and the coordination numbers CN)
CN) as a function of
x, for Nd
Ndoo-x
90 - x Fe x AI\o
x AI ,o melt-spun ribbons, at room temperature.
Thickness Fe-Fe Fe-NdCAD Nd-Nd
x
tC~m)
tClJm) rCA) N rCA) N rCA) N
25 2.48 4.45 2.77 4.53 3.35 3.99
35
120 2.62 7.81 2.87 4.73 4.13 3.59
25 2.54
254 8.65 2.85 4.85 3.44 5.72
40
120 2.51 7.69 2.76 7.39 3.38
338 8.30
25 2.53 7.54
~.54
, 2.80 5.93 3.29 3.85
50
120 2.43 6.46 2.71 7.41 3.35
335 4.55
332 H. Chiriac and N. Lupu

Let's consider atoms Iike like rigid dense spheres randomly packed in a given
structure. If we ignore the intensity of the chemical bonding, thus the
interatomic distances are equal with the sum of the atomic radii. radi i. Considering
= 1. 82
r Nd = 1.82 A, r Fe =
= 1.27 A and r AI = 1.43
= 1. 43 A, the theoretical values for Nd-Nd,
Nd-Fe, Nd-AI, Fe-Fe, Fe-AI and AI-AI atomic pairs are respectively: 3. 64,
3. 09, 3. 25, 2. 54, 2. 7 and 2. 86 A. Comparing these values with those
3.09,
calculated from experimental data, one observes that Fe-Fe values fit very
well, whereas the others are larger than experimental ones. This can be
explained in the hypothesis of the existence of randomly distributed Fe clusters,
while Nd and AI atoms are distributed randomly between these clusters. We will see
in the next section the influence of these structures on the magnetic behavior of
the Ndgo-xFexAllo-ySiy amorphous alloys.

9.4.4 Magnetic Properties of N~o-xFexAlIO-ySiy


Nd90-xFexAlIO-ySiy Bulk Amorphous Alloys
9.4.4. 1 Composition and Thickness Dependence
The structural investigations on Ndgo - x Fe x A1 1o - y Siy amorphous melt-spun
Al lo
ribbons and cast rods revealed a new type of disordered structure consisting of
a dense random packing of nanometer-sized atomic clusters, whose size is
dependent on the samples thickness, i. e., the preparation conditions.
Consequently, one expects complex magnetic to be behavior dependent on
thickness and a composition.
Values of the magnetization measured in a field of 800 kA/m and a
coercive field obtained at room temperature for Ndso Fe40 AI lo amorphous
Al IO
samples are given in Table 9. 7. The large values of the coercive field can be
explained by assuming the existence of a microstructure consisting of very small
magnetic clusters, whose size approaches a single magnetic domain, dispersed in
the amorphous matrix, in agreement with the neutron diffraction results as well as
with the previous results obtained for melt spun Nd-Fe amorphous alloys
(Croat, 1982). The size and the composition of these magnetic clusters are
strongly dependent on the cooling rate. The sizes of these magnetic clusters
are small enough and cannot be evidenced by XRD measurements.
Table 9.7 Magnetic characteristics of Nd50
50 Fe,o
Fe40 Al lO amorphous ribbons and rods, at room
temperature (Chiriac et al. , 2001 d) .
Sample Thickness(mm) /..10 M(1o
fJo M(lo kO.)
kOe) (T) i H, (kA!m)
Amorphous ribbon 0.030 033 64
Amorphous ribbon O. 055 0.24
024 123
Amorphous ribbon O. 085
0.085 025 226
Amorphous ribbon O. 105 o 32
032 274
Amorphous ribbon O. 142 023 286
Amorphous rod O. 6 025 244
Amorphous rod 1. 8 024 246
Bulk Amorphous Magnetic Materials 333

The higher values obtained for the coercive field of the thick amorphous
ribbons comparatively with thin ones are related to the more relaxed
microstructure developed in the first ones. It is worthwhile to note that the
magnetization for the ribbon of 105 IJm j..lm is almost the same as for the ribbon
30 IJm,
j..lm, whereas the coercive field is 4 times larger. This special behavior can
be explained by the largest degree of structural disorder existent in the ribbon
of 105 IJm,
j..lm, in agreement with the largest crystall ization energy obtained from
crystallization
DSC curves. In these conditions, an external appl ied field causes the
alignment of more magnetic domains relative to its direction.
The microstructure is more homogenous in amorphous rods due to their
dimensions and lower cooling rates, and consequently the values obtained for
coercive field are smaller than for the thick ribbons.
Ndgo-xFexAllo (x = 35 - 50) amorphous alloys exhibit large coercive fields
Ndgo-xFexAI1O(x=35-50)
at room temperature, even for low applied fields, regardless of their thickness
and preparation techniques. In Fig. 9. 17 the hysteresis loops as a function of a
the Fe content for 3 different kinds of Ndgo - x Fe x Al lO lo samples are pres~ted
comparatively: fully amorphous thin ribbons 25 IJm j..lm in thickness (a), fully
amorphous (x = = 35, 40) and for comparison, partially vitrified (x = = 50) thick
0.5 0.5
0.4 Ribbons 25 f..lm
f1m x 2 mm _- __ 0.4
0.3 I1;'-
" """,,,,,,," ".'" 0.3
0.2 .J",i::'~~
J__ i:~~ 0.2
f- 0.1 f- 0.1
'-'
~ 0 ~
~
00
l--D.l
l--D.I -x=35 :: --D.l
l--D.I
-{).2
--D. 2 ---- x=40 --D.2
--D. 2
x=45 ....
--,- x=50
--D.3
-{).3 ,,' ' ~,_.", I --- x=50
-{).3
--D.3
--D.4 __ - - -'- --D.4
--D.5 --600
--{i00 -400 -200 0 200 400 600 --D.5--{i00 -400 -200 0 200 400
--D.5--600 600
Applied field (kA/m) Applied field (kA/m)
(a) (b)
0.5 II
04
0.4 mm~
Rodsl1 mm
Rods
. I "' ~:.:::',
:,."...

~ 1; ._d87
0.3

-z{f:&f~5~
u

0.2
f- 0.1

l--D.~ -/LiP---=-x~35-
'-' 0
l--D.I
--D.2 .----_.,: ../.- _1-
--:::~:::::::'L(:':;;:' .:-~-: :~~
---- x=40
--D.33 ..,.:: .. - - I
--D. x=45
--D.4
--D.4 ... x=50
--D.5--{i00
--D. 5--600 -400 -200 0 200 400 600
Applied field (kA/m)
(c)

Figure 9.17 M-H hysteresis loops of Ndgo - x Fe x Al,o melt-spun ribbons 25 and
90 -

respectively 120 IJm in thickness and cast rods 1 mm in diameter, at room temperature
CLupu et al. , 2M
CLl1pu 1b) .
2001b).
334 H. Chiriac and N. Lupu

ribbons with a thickness of about 120 IJm (b) and fully amorphous cast rods
1 mm in diameter (c). The magnetization increases about 2 times by
increasing the Fe content from 35 to 50 at. %, regardless of the amorphous
sample's shape, the same behavior being observed for high fields up to 9 T.
The coercive field presents a strong dependence on the composition. One
observes that the on set of crystall ization in the Nd40 Feso Al lO thick ribbon
crystallization
results in the decrease of the magnetization, while the coercive field has
almost the same value as that obtained for fully amorphous thick ribbons with a
smaller content of Fe because of the existence of the amorphous residual
phase. Thus, the large coercive fields of Nd-Fe-based ternary bulk amorphous
alloys are related to the existence of the amorphous phase, its disappearance
leading to the drastic decrease of the coercive field and moreover, to the
disappearance of the ferromagnetic properties, as we will see in the
following. The decrease of the Fe content results in the decrease of the
coercive field of about 2 times for thin amorphous ribbons and in its increase of
about 2 times for amorphous cast rods. For thick amorphous ribbons the
coercive field is almost the same, regardless of the Fe content. For higher
external fields (larger than 3 T) the coercive field does not change significantly
with the Fe content for each different type of ribbon or for cast rods. The
difference in the absolute value of the coercive field as a function of the
amorphous samples' thickness results from the volume rate between Fe-based
magnetic clusters and the homogenous Nd-rich matrix through the cooling rate.
9. 18 shows the variation of magnetization measured at 1260 kA/
Figure 9.18 kA/m m
and coercive field, as a function of y, for NdsoFe4oAl1O-ySiy melt-spun ribbons
and cast rods. The increase of the thickness of the amorphous ribbons leads to
the development of more relaxed structures in the as-cast state and the degree
of structural disorder is diminished. Thus, the magnetization of the thick
amorphous ribbons 85 IJm thickness decreases in comparison with the value
obtained for thin amorphous ribbons, while the coercive field increases with
the increase of the amorphous samples' thickness. While the magnetization of
the amorphous ribbons does not change significantly from replacing AI with Si,
the coercive field increases. This behavior due to the different influence of AI
and Si on the structure and magnetic properties of Nd so Fe40 Al lO - ySiy alloys,
NdsoFe4oAI,o-ySiy
even if they have relatively narrow atomic radii (r = 1.43 A; rSi
AI =
(rAI = 1.32 A),
r Si = A) ,
is in agreement with the results reported for Nd-Fe-AI and Nd-Fe-Si crystalline
al. , 1995; Allemand et al.
alloys (Le Breton et a!. a!. , 1990). It is worth noting that
the ferromagnetic properties of Ndso Fe40 A1 1o -- y Siy alloys are related to the
Al lO
existence of the amorphous phase. Diminishing the amount of the amorphous
phase in the as-cast state, the magnetization and the coercive field decrease,
as it can be observed for samples containing more than 5 at. % Si (Fig. 9. 18
and Table 9. 8). The hypothesis of the existence of small magnetic clusters
embedded in the amorphous matrix, whose composition and size are very
sensitive to the cooling rate and the preparation process, is confirmed by the
improvement of the remanence and coercive field by applying suitable
Bulk Amorphous Magnetic Materials 335

annealing on the molten alloy before ejection (Chiriac et al. , 2001e).


2001e) .
0.35~----------~
0.3S~----------~
300
0.30~
0.25 ;\\
0.2S;x:-....
-.------.T---.
.....-----... ..
T---~ ~.. 6
:::..,... 6""0'
""(j --<)-
--<)-
....- . ;
___ '-'0-
-0 250

P 0.20
---
"-'
.....
.~ i't'1't
.... t'
t1 ---.... - -0-
_ 0..
0- - - .....
-< ,,,,:
'.'7
""/
... .....
..... 200
E
1~ 0.15 ---is\
i!i\ -Ribbon
--Ribbon 30 lJ.m f!m '\.... 150 :;;:
ISO:;;:
6c
......'
.....' -o-Ribbon 8SlJ.m
-C.-Ribbon 85 f!m
0.10 "'> , . .
...:,- ....t.....
t ;.... Rod 0.6 0 6 mm$ mm~ \
\. "
100:X::
......::-- _.".-
....:..\. _.'(J.- Rod 1.8 1:8 mm$ mm~
\
\. 50
0.05
.....~. :-- "

.~'-.~..
:--- .Jl. - - - -'
oOt:.-_-'-_-.L----::L;~_.L:::.:.::!t;::.:.::.:J::.:.::.:.::.:.=
t:.-_-'-_-L----"~:... ..._....L..:.::.:
... :.<e.t..::.:
...:::J
.. ::.:
...::.: ...::.:.:J
...::.: .... ~'-.- <
..... :::::........ 0
o 2 4 6 8 10 0 2 4 6 8 10
y(at.%Si) y(at.%Si)

Figure 9.
9, 18 The magnetization at 1260 kA/m and the coercive field as a function
of Si content for Ndso Fe,o
Fe40 A1
Alto-
1o - y Si y melt-spun ribbons and cast rods (Chiriac and
Lupu, 2001a).

Table 9.8 Magnetic characteristics of Ndso Fe,o


Fe40 AI
Al lOIO melt-spun ribbons and cast rods at room
temperature CChiriac et al. , 2001b).
2001 b).

Thickness {JoM
/JoM ,He
Alloy Sample Structure
(mm) (T) (kA/m)

Melt-spun ribbon 0.030 0.25 160 Amorphous


Melt-spun ribbon 0.045 0.20 248 Amorphous
Melt-spun ribbon 0.090 0.10 305 Amorphous
Ndso
Nd Fe40 Sho
so Fe,o Melt-spun ribbon 0.115 0.046 290 Crystalline
Melt-spun ribbon 0.142 0.021 56 Crystalline
Cast rod 0.6 0.005 4 Crystalline
Cast rod 1.8 0.001 4 Crystalline

9.4.4.2 Magnetic Transitions in the Temperature Range 4-300 K


In order to obtain information on magnetic coupling of the hard magnetic phase
detected in Nd90 Al ,o
go -- x Fe x AI lO -- y
y Siy bulk amorphous alloys, the low temperature

dependence of the magnetization curves and B-H S-H hysteresis loops was carried
out for Nd50 so Fe40 AIAl ,o
lO amorphous samples with different thicknesses
(Fig. 9. 19). A different behavior is observed for ribbons and rods. While at
reduced temperatures the ribbons have no B-H S-H hysteresis loops, increase of
the temperature results in an increase of the magnetization and coercive field.
The increase of the temperature above 200 K results in the decrease of the
coercive field. For amorphous rods there is hysteresis even at reduced
temperatures, and the coercive field increases with increasing the temperature
up to 200 K and decreases thereafter. The coercivity at low temperatures is
probably larger, but the available fields (800 kA/m) are not enough to
336 H. Chiriac and N. Lupu

saturate the sample and therefore only minor loops can be obtained at low
temperatures. This can be the reason for the observed displacement of the B-H
loops in the second quadrant. Also at low temperatures the antiferromagnetic
contribution of Nd is more pronounced, this contribution diminishes with the
increase of the temperature and as a consequence above 200 K the B-H loops
of NdsoFe4oAlIO
Nd50Fe4oAlIO bulk amorphous alloys are centered. The maximum appearing
(150-200
in the coercivity over a narrow range (150 - 200 K) suggests the existence of
one pronounced anisotropy.
0.4
Ribbon-30 ~m

0.2

E o0
i -5K
----lOOK
-0.2 .. 150K
150K
-~200K
- ' 200K
--300K

-0.4
-{).4 L....-'_--'-_-'-_-'-_-'--_-'-----"'----'-_-'-_
L-.l_--'-_-'--_-'-_--'--_-'---"_--'-_-'--_
-800 -600 -400 -200 0 200 400 600 800
Applied field (kA/m)
(kAhn)
(a)
0.3
Rod-I.88mm~
Rod-l mm~

0.2

0.1

F='
f=' 0

~ -0.1
-0.2

-0.3 '-:::-:':-::--::-:-:,...-=---=-':-:--:--=,...-:::-:----::c'c::--::-':--=-'
-0.3 L-.l_---L_-L-_--'--_-'---"_--'-_---'---_---L-.J
-800 -600 -400 -200 0 200 400 600 800
Applied field (kAhn)
(kA/m)
(b)

Figure 9. 19 Magnetization as a function of applied field and temperature


dependence of B-H hysteresis loops of Ndso Al ,o
50 Fe40 AI lo amorphous ribbons and rods
(Chiriac
CChiriac et al. , 2001 d) .

The difference in the absolute value of the coercivity as a function of the


samples' thickness may result from the volume ratio between the magnetic
Bulk Amorphous Magnetic Materials 337

clusters phase and the homogenous amorphous matrix through the cooling rate.
However, at higher temperatures the coercivity is reduced so that the
available magnetic field causes more magnetic domains to participate in the
magnetization reversal. This can be observed also from magnetization curves
plotted at temperatures above 200 K. Irreversible processes dominate the
initial part of the magnetization curves at low fields, which indicates that there
is no free displacements of domain walls. Therefore we can assume that the
magnetic clusters form an assembly of isotropically distributed uniaxial single-
domain particles.
The temperature dependence of the magnetic characteristics at low
temperatures indicates the presence of more than one magnetic phase below
150 - 200 K, depending on the cooling
cool ing rate and the thickness of the amorphous
samples. This is in agreement with the multiphase transitions inferred from the
low temperature ac-susceptibility measurements, as it can be seen in
Fig. 9.20. A slight dependence on temperature appears with the increase of
the sample's thickness and probably is due by the changes of the
microstructure depending on the cooling rate. The imaginary part (out-of- Cout-of-
phase ac-susceptibility X") clearly shows differences between ribbons and
bulk amorphous rod being related to the more relaxed amorphous structure
X' (in-phase ac-susceptibility) at around
existent in bulk samples. The peak in x'Cin-phase
10K seems to be the Neel temperature of the main Nd-rich antiferromagnetic
phase or the freezing temperature below which Nd magnetic moments will be
locked to their local easy axis. It is interesting to note al50
al&o the existence of
one step on the real part around 100 K, just before X' starts to increase,
which corresponds to the pronounced increase of the coercive field obtained by
dc-magnetic measurements.

5.0
4.0
- Ribbon 30 11m
~m
~
::;;) 4.0 ~m
Ribbon 85 11m
'1E: --- Ribbon 14211m
142 ~m
~ 3.0 - - Rodrp=1.8
RodiJ>=1.8 mm
;;:::
;::::: 3.0
bfJ
bIJ

'1Eu: "
:'
b
::l
E
~ 2.0

''
2S
~ 2.0
'"
'I'
b
o
~ 1.0
-?-<
-?<
,,~_
"._~_._.. _.. _.. _.':~
_.....
/
0"-
O\..

---
ao 50 100 150-WO-
150-WO./
Temperature (K) ~

' ... '.-,~ .,, ....... ".' '.""


'.., ',' :-.',
:-,',
.."--:'~""""'"
~ . ',-.':
:-: ...',',-"-
~: "':'
'.-,': ~:"':' -.
,.... ... ,:",,,:,,',' ':""".'.
...., ... " .. ~
.....
:-.. ,:._""'l"'I"~"''''''

ao 50 100 150 200


Temperature (K)
(X') and imaginary part (X")
Figure 9.20 Real part <X') <X") (inset)
<inset) of ac-susceptibility
at 0.8 kA/m and 182 Hz for Ndso Fe.o
Fe40 Al IO melt-spun amorphous ribbons and cast
AI lO
<Chiriac et al. , 2001 dl
rods (Chiriac d) .,
338 H. Chiriac and N. Lupu

Figure 9.21 presents the dependence on temperature of the magnetization


and coercive field for Nd90 go - x Fe x Al IO thin amorphous ribbons, thick amorphous
AI lO
or partially vitrified ribbons, and amorphous cast rods. The humps appearing
in the coercive field over a narrow range (150 - 220 K) for amorphous ribbons
containing a large amount of Nd (over 50 at. %) and for cast rods in the entire
range of compositions suggest the existence of one pronounced anisotropy of
the magnetic clusters. The decrease of the coercive field with the temperature
is similar to that obtained for similar crystalline compounds and is believed to
1.2 500
Ribbons 25 flmX2
J.lm x 2 mm Ribbons 25 flm
J.lm x 2 mm 0
1.0 _~_.....".._--.......t:,.
-,.,.....-~-----......fj. :FE=< 400
~
6
0.8 ...... x=35
--- x=35
...... x=35 ~ 300
~ --- x=35 t;::
t;:
-0- x=40
~ 0.6 -0- x=40 <l) -<:r- x=50
-fr-
-<:r- x=50
-fr- .~ 200
.2: 200
:: 0.4 2
<l)
o
U
0.2

o 50 100 150 200 250 300 100 150 200 250


Temperature (K) Temperature (K)
(a) (b)
0.35
Ribbons 120 l.lInX4
flm x4 mm 600 J.lm X
Ribbons 120 l-un X 4 mm 0
0.30
~ 500
E 500
0.25 =<
6~400
400
P
~ 0.20 -0
"0
Q)

~ 0.15 " 300


t;:
t;::
<l)
>
300

0.10 '2
.~ 200
:go
<l)

u
U

o 50 100 150 200 250 300 100 150 200 250 300
Temperature (K) Temperature (K)
(e)
(c) (d)
0.35
Rods I mm 600 Rods 1I mm
0.30
~ 500
E 500

~400
~400
-0
"0
Q)
"
t;:
t;::
<l)
>
300

...... '2
.~ 200
--- x=35
x=35 <l)
o
-0- x=40 U
0.05
-fr- x=50
-<:r-
o ~
50 100 I~
150 200 2~
WO 250 300 o 50 100 150 200 250 300
Temperature (K) Temperature (K)
(e) (I)
(f)

Figure 9.21 Temperature dependence of the magnetization at 1 T and coercive field for
Ndgo - x Fe x AI IO amorphous ribbons and cast rods as a function of the Fe content (Lupu
et ai, 2001b).
Bulk Amorphous Magnetic Materials 339

be due to a change of the size of the magnetic clusters domains, which


probably approaches to the superparamagnetic regime. Due to the more
homogenous structure existent in the thin ribbons the clusters' size is smaller
and consequently the interactions between them are not so strong. The
behavior is typical for a spin-glass system with random configuration of the
magnetic spins, in which thermal activation energy is enough to destroy the
magnetic order. The coexistence of two types of magnetic order: long-range
ferromagnetic order which gives the macroscopic behavior, and short range

5.0

4.0
~
M 3.0
4.0 ~

----
J1111
Ribbon 25 pill
.-,- Ribb.onl20
Ribbon 120 . .n.
Rod</F
RodlF 1 mm ,.
1111111
/ry/
/
~lIn/-~'v
~.,
,~
.
/ /

~
'~

II/
c:
~
00::
~ 2.0
3.0 :(
7
0

x
2.0

1.0
/
/
#/
o "",
/" ,.;7
x 2.0 ... ~.._.~/~.-
';.<
o 50 100 150/""
r em perature (K L/:
;>9

1.
0
f:.':: L
r

6.0

5.0

bii 4.0
""
E
:;:;:
E 3.0

I
0
x 2.0
';.<

1.0

o0 50 100 150 200 250


Temperature (K)

Figure 9.22 Real part CX') and imaginary part CXU) Cinset) of the ac-
susceptibility for Ndgo-xFexAllOCx=35, 50) melt-spun ribbons and cast rods.
340 H. Chiriac and N. Lupu

ordered structures which determines the large values of the coercive fields in
Nd-Fe-based bulk amorphous alloys is proved by the differences between zero-
field-cooled and field-cooled magnetization curves (see next section). From
the temperature dependence of the coercive field for Nd 40 Feso Fe50 AI
Al 10
lo thin
amorphous ribbon 25 IJm !Jm in thickness one observes the existence of two
maximum for the coercive field: one around 200 K which is similar to that
obtained for the other bulk amorphous samples and the new one around 25 K,
whose origin is unknown, but probably is caused by the larger amount of Fe
which determines a more homogenous topological and magnetic structure. It is
worth noting that the coercive field for this ribbon in the as-cast state is the
Ndgo-xFexAIIO amorphous alloys with x = 20 at. %-
smallest one measured for Ndgo-xFexAllO
60 at. %. The magnetization increases with the temperature increase, this
increase being more pronounced for thin ribbons 25 IJm \Jm in thickness containing
the largest amount of Fe in composition.
In agreement with DC-magnetic results, the real part of the ac-
susceptibility (X') (Fig. 9.22)
922) shows a second peak at around 70 K, which
appears also for the imaginary part (X"), and is more pronounced for thick
ribbons containing 50 at. % Fe. Above 70 K, the Nd-rich matrix and Fe-based
magnetic clusters are magnetically uncoupled and the response of the system is
weakly ferromagnetic. The strong increase of the susceptibility with the
temperature above 70 K represents the response of the ferromagnetic clusters
that cannot be saturated by the available applied fields.
fields
The second maximum on the AC-susceptibility curves is observed also for Nd50so
Fe40 SilO melt-spun ribbons and cast rods (Fig. 9.23), being more pronounced
14
2.5 !
- Ribbon 30 flm
pm
12 ~ 2.0
SJ
"1:
r~~ Ribbon 85 11m
fllll
f.
, ~ .... pl11
- ... Ribbon 142 11m
": 1.5
10 I.
l. :.
'
:
..
~ "Eu - - Rod=I.8 mm
Rod=1.8mrn
~
,. ".
/. 0 1.0
~
}O ..-
'-" .1 xv
8.0 ''><,., 0.5
"'
~
p
7.~ 6.0 0 50 100 ISO
I50 200
o Temperature (K)
(KJ
xX
~

-",
";-< 4.0

2.0

o 50 100
j IP.
~
150
'i-"'l"';
"!-,"'.'-...,',' 'T"'T Ii .....
'F ..J"tTi-..,.'/.,..., r I ','l""'". ',' "'T"','!' ~,' .... , ...
1"',' 11"",',"",' IT'1F.'....

200
.. ,'

Temperature (K)

Figure 9.23 Real part (X') and imaginary part C X") (inset) of ac-
(XU)
susceptibility at 0.8 kA/m and 182 Hz for Nd
NdsoFe4oSi,o
50 Fe40 SilO melt-spun ribbons and
cast rods (Lupu al , 2001b)_
CLupu et aI., 2001b)
Bulk Amorphous Magnetic Materials 341

for samples which consist in a mixture of amorphous and crystalline phases.


9.4.4.3 Thermal and Theromagnetic Treatments-the Influence on the
Magnetic Properties
The hysteresis loop at room temperature is the response of the magnetic
clusters phase, because the Nd-rich phase is paramagnetic and has no
contribution. The annealing at temperatures lower than crystallization
~ 773 K).
temperature ( T x """ K), as shown in Fig. 9. 24, leaves the magnetic
properties almost unchanged, which we suggest is due to the thermal stability
of the amorphous alloys. The onset of crystallization
crystall ization leads to the metastable
phases formation, which have larger coercive fields but smaller magnetizations
in comparison with the amorphous phase. These metastable phases disappear
after annealing at 773 K, above this temperature the sample becomes
paramagnetic.
0.4

- - as-cast
- - - 10/373 K
0.2 10/423 K
-,-,10/473
-_.- 10/473 K
."."",,,,,:.::"''''

E 0
------
---- ------
.. -----
1~ -. 10/623 K
.... 10/673 KK
-0.2
10/723 K
_._-- 101773K
---- 10/773 K

-0.4 '---'-_---'_
'-------'-_ _L-_-----'-_
_-'--__---'---_
_L-_----'--_
_L - _ _-'-----_-----"-_
- - ' -_ _- ' - _
-600 -400 -200 0 200 400 600
Applied field (kA/m)

Figure 9.24 Hysteresis loops as a function of annealing temperature for Ndso Fe40 Al
50 Fe,o AI IO
lO
amorphous ribbon 30 IJm in thickness (Chiriac and Lupu, 20011).

The coexistence of two types of magnetic order, short-range spin-glass-


like order and long-range ferromagnetic order even in amorphous rbds, is
Iike
evidenced by zero-field and field cooled thermomagnetic measurements in
Fig. 9.25.
The presence of the maximum on the M ZFC curves and the bifurcation
between MzFcand M Fc FC curves prove the coexistence of two magnetic phases.
The displacement of the M ZFC maximum towards high temperatures with the
increase of the thickness of the amorphous samples is determined by the fact
that the thinner the ribbons, the higher the glassy disorder and the higher the
thermal energy required to destroy it. The cusp in the M ZFC magnetisation
curves (Fig. 9.24), whose position changes with decreasing temperature by
increasing the magnitude of the external field, and the bifurcation of the M zFc
ZFC
342 H. Chiriac and N. Lupu

0.16

0.12

MZFCMFC
E
Eo.08
0.08 ~ - .--
-_.- Ribbon 30 J.lm
J-lm
1 ~ -- <_.
-
- -+
~_.
- ' --
- -- ..0_
Ribbon 85 J.lm
J-lm
-<>_.- Ribbon 142 J.lm
J-lm
0.04 ~ -.,_. Rod-</>= 1.8J.lm
-..,. 1.8J-lm

o
400 450 500 550
Temperature (K)
(a)
0.20

0.16

,
0.12 ,, MZFCMFC
p
E ~ - -- . Ribbon 30J.lm
---. 30 J-lm
1 0.08
lO.08 ---+-
-

-
+ - - ..,-

- -+
<-.-

- ' --
- - ..0_
Ribbon 50 J.lm
J-lm
Ri bbon 85 J.lm
-<>- Ribbon J-l m

0.04
,
O~;:----:::-;:-,;:-----~~~=::;;=:~==~
O~:c---co-:':::-~~~==;;;=:::::;;:='=~
300 350 400 450 500 550
Temperature (K)
(b)

Figure 9. 25 M zFc
ZFC and M Fc
FC magnetization curves of Ndso Fe,o Al,o
Al lO and Nd so
(He,! = 40 kA/m)
Fe,o SilO amorphous samples with different thicknesses (He,'
(Lupu and Chiriaco
Chiriac, 2001a).

and M Fc curves are characteristic to systems in which ferromagnetic long-


range order and spin-glass or cluster-glass short range order exists (Fisher and
Hertz, 1993). The same changes were observed for both Nd50 Fe40 Al lO
amorphous ribbons and rods, the only differences being observed for the
position of the cusp as a function of the amorphous ribbons' thickness and
being related to the thermal effect.

9.4.5 Mossbauer Spectra


The Mossbauer
M6ssbauer spectra in Fig. 9. 26 have been fit with two overlapping
sextets, one of which is attributed to the hard phase and another one to dilute
Fe in the Nd-rich phase. The hyperfine parameters obtained for the best
spectral fits are given in Table 9.9. Assuming that the iron magnetic moment
/-IFe is proportional to the hyperfine field with a constant of proportionality of
IJFe
Bulk Amorphous Magnetic Materials 343

15 TIPe (Qi et aI., 1992), the average Fe moment in NdsoFe4oAI,o-ySiy


T/,uB (Oi NdsoFe4oAllo-ySiy is
only 1. 25 - 2. 20,uB.
1.25 20Pe. This can be ascribed to the strong hybridization of the AI
or Si electrons with the iron 3d bands (Hu et aI.,
al. , 1992) or to the disturbance
of the parallel alignment of the spins by local anisotropy.

I00. 0 ~1-~ir:i"~W
100.0 ~ir:i-iIi04W

99.8

~c
t:
99.6
o
.<ii
'Vi
'
tI'l
.~
99.4
99.4
'"
~ 99.2
,,
99.0 y=IO
98.8 L- -
----
-----::'L
-_ LL
-_ . --
_ .L
_ - '.--:_
-_.L
-. -:
'- _--
_ '-
._ . .-L
.L _---L
_--_
' :- '_
:-..
_ ..
-J' ,,-------
--:'
-9 -7 -5 -3 -I 1I 3 5 7 9
Velocity (mm/s)
(a)
.- ...-.
.-.
I00.0 ~,~~ML
100.0 ~,~~e.wL ::. ...... -
99.8

~ 99.6
ct:
o
.~ 99.4
s'E
~~ 99.2
99.2
~
f-
99.0

98.8 y=O
98.6 L -_L.-_L.-_.L.-_-'-_-'-_..L--_-':-_-':----:'
'----..1...---'----'------'------'-------'-------'--'---------'
-9 -7 -5 -3 -I 1I 3 5 7 9
Velocity (mm/s)
(b)

Figure 9.26 Mossbauer


M6ssbauer spectra of Ndso Fe40 A1
AI,o-
1o - x Si
Siyy amorphous alloys at room
temperature (Chiriac
CChiriac et al. , 2001b).

Table 9.9 Hyperfine parameters for Ndso Fe40 A1 10 - ySiy


Aho- CY = a
ySi y (Y 0,10)
10) amorphous alloys (Chiriac
CChiriac
et al. , 2001b).
Isomer shift Quadrupole shift
ShfCT)
BhleT) Area
(mm/s)
Cmm/s) (mm/s)
Cmm/s)
18.82 -0.09
-0. 09 0.211 91
y=O
29.18 0.05 0.63 9
21.18 -0.12 0.224 92
y= 10
33.05 0.00 0.57 8

The corresponding 2 magnetic phases hyperfine fields are presented in


344 H. Chiriac and N. Lupu

Fig. 9. 27, as a function of y. One observes that the hyperfine field of the
phase in proportion of 90 % is much smaller than the hyperfine field of the 10%
phase. Due to the proportional ity between hyperfine field and Fe magnetic
moment. we can conclude that the 10% phase is the magnetic clusters phase,
moment,
whereas the 90 % phase corresponds to Fe atoms dispersed in the Nd-based
matrix.
35

30

E
~
25

20
0 2
Phase 1(90%)

4
I(90%)

6
AI (al.%)
(at.%)
----
I

8
-----'
- --'

Figure 9. 27 The hyperfine field dependence on AI content for Nd50


10

so Fe,o AI,o- y Sly


Si y
melt-spun amorphous ribbons containing two magnetic phases.

9. 5 Relationship between Microstructure and Magnetic


Properties in Bulk Amorphous Alloys

9.5.1 Soft Magnetic Bulk Amorphous Alloys

The improvement of the soft magnetic properties of the Fes6


Fe56 COl Nil Zr6 M1 5Nb
S Nb2 5820
2 s 8 20
(M = Ti, Ta, Mo) bulk amorphous alloys is achieved by the choice of the
optimal composition, by the increase of the amorphous samples, thickness or
by thermal and thermomagnetic treatments applied at temperatures below
Thus, the softness of the Fe-based bulk
glass-transition temperature (T g), Thus.
amorphous alloys is closely related to the existence of one relaxed disordered
structure.
The composition leads to the achievement of a higher degree of
randomness in the amorphous state due to the different atomic sizes of the
components. This condition is necessary also to be satisfied in order to
enlarge the glass-forming ability.
The increase of the amorphous samples,
samples. thickness results in a more
relaxed amorphous structure in Fe-based bulk amorphous alloys. Moreover. Moreover,
the preparation method seems to play an important role in the obtaining of one
more relaxed structure. For example.
example, the cool ing rate decreases from 10 lO 5s-
Bulk Amorphous Magnetic Materials 345

10 6 K/s tJm to 10 2 -10


K/ s for thin amorphous ribbons 25 IJm 3
- 10 K/s
K/ s for thick amorphous
2
ribbons (thicker than 100 IJm)
tJm) and to 10- 10 K/ s for cast samples (rods or
rings). Additionally, the structure of the cast amorphous rods is free of
additional stresses induced during the preparation of the ribbons by the melt-
spinning method. It is worth noting that the soft magnetic characteristics of the
rods are similar to those of the rings; we are even expecting to have some
circular magnetic domains in the last ones. The understanding of this behavior
needs more careful investigation.
Thermal or thermomagnetic treatments applied at temperatures below
crystallization temperature results in the improvement of the softness of the
amorphous melt-spun ribbons, i. e., the increase of the magnetic
permeability, the decrease of the coercive field and core losses, through
structural relaxation processes. By anneal
annealing,
ing , the amorphous structure is
relaxed mainly by the diminishing of the residual stresses induced during the
preparation.
It is worth observing that the soft magnetism in Fe-Zr-B-based bulk
amorphous alloys is related just to the existence of the amorphous phase. The
on set of the crystallization results in the diminishing of the magnetic
permeability-
permeabil ity - the increase of the coercive field and magnetization - thus the
material's softness.

9.5.2 High-Coercivity Bulk Amorphous Alloys


Structural, magnetic (dc and ac) and M6ssbauer spectroscopy studies on
Ndgo-xFe
Nd go - . Fe.x Al 1o
lO -
- y Siy melt-spun ribbons and cast rods show that these bulk

amorphous alloys are structurally"


structurally "glassy"
glassy" but magnetically"
magnetically "granular".
granular". The
granular magnetic structure consists in Fe-based nanosized clusters dispersed
in amorphous Nd-rich matrix. Due to their very small dimensions they cannot
be evidenced by X-ray diffraction measurements, but their presence is proved
by the specific magnetic behavior of these bulk amorphous alloys, especially
by the high coercive fields in the as-quenched state and the dependence on the
cooling rate. The composition and size of the magnetic clusters as well as the
composition of the matrix in which the clusters are embedded are very
sensitive to the preparation conditions, mainly the cooling cool ing rate and the thermal
history of the molten alloy.
Fe-based magnetic clusters are coupled between them through the
amorphous matrix and the exchange coupling is strongly influenced by the
magnitude of the Nd3+ single ion local anisotropy. Whereas at low
temperatures the anisotropy plays the predominant role in the macroscopic
magnetic response of the system, the increase of the temperature diminishes
its effect and the' the main role is attributed to the ferromagnetic exchange energy.
At low temperature, the local anisotropies of the Nd3+ ions are very strong and
oriented randomly, thus they can "freeze" randomly the magnetic moments of
Fe and Nd. The anisotropy axis corresponding to the Fe-Nd pairs are oriented
346 H,
H. Chiriac and N. Lupu

randomly, and a small external field is not enough to al align


ign them.
The ac-susceptibility measurements as well as the variation of the
coercive field on temperature in the range 4 - 300 K indicate magnetic
transitions that suggest different local arrangements of the Nd atoms, similar to
dhcp (double hexagonal compact) and fcc crystalline phases. The first
transition appears around 10K and seems to be related to the Neel transition
temperature of the Nd antiferromagnetic. Another magnetic transition is
observed around 70 K and is due to the transition from a completely disordered
magnetic phase to a sperrimagnetic non-coil
non-collinear
inear one.
Nd in the dhcp form is antiferromagnetic and exhibits two transition
temperatures: the first one at 7. 5 K represents the ordering temperature for
the cubic Band C sites, and the second one at 19.9 K is the ordering
temperature of the hexagonal A sites (Moorjani and Coey, 1984). But,
depending on the nature and the size of the nearest neighbors, Nd could also
exist in fcc allotropic form with ferromagnetism below 29 K. Thus, Nd could
exist either in dhcp or fcc form in the amorphous matrix of the Nd-Fe-based
bulk amorphous alloys. Fe also could be ferromagnetic (r > 1. 5 A) or
anti ferromagnetic (r< 1.5
antiferromagnetic 1. 5 A) depending on the coordination number and the
nature of the nearest neighbors (Bozorth, 1993).
Different magnetic structures can be obtained depending on temperature
and the predominant coupling between Fe and Nd atoms. This is probably the
explanation for the lack of the hysteresis loop at temperatures between 10 and
100 K and low external fields 1 n. T). In this range of temperatures Nd-rich
matrix and Fe-based magnetic clusters are completely uncoupled, but the
exchange coupling
coupl ing within the clusters exist. If, for any reason, the coupling
coupl ing
between clusters is not possible, the macroscopic response of the system is
weakly ferromagnetic.
The maximum of the coercive field is attained at that temperature at which
the anisotropy energy and the exchange energy have similar values, i. e. , the
magnetic clusters have the size of the single magnetic domain (Cullity, 1972).
Taking into consideration the critical size for the Fe single domain of about
1000 A and the strength of the local anisotropy of the Nd3+ ions results that the
Fe magnetic moment is not aligned on a direction but forms a cone with an
angle of about 30 (Moorjani and Coey, 1984). Thus, the magnetic structures
are non-collinear and the critical size of the single magnetic domain is smaller
than 100 nm.
In Nd-Fe-based thin amorphous ribbons, due to their reduced dimensions,
high cooling rate (which does not allow the structural relaxation) and thermal
effects, the system tends to a superparamagnetic regime at room
temperature, and consequently the coercive filed is not very large. It is known
that in materials containing single magnetic domains the magnetization is due
mainly to the magnetic moments rotations. If'the value of the applied field is
not large enough, the measured hysteresis loops are minor loops, and the
measured magnetization is far from its saturation. This effect is more
pronounced in materials in which the magnetic anisotropies play an important
Bulk Amorphous Magnetic Materials 347

role in magnetization processes. Nd-Fe-based bulk amorphous materials with


3
strong local random magnetic anisotropy of the Nd Nd3++ ions, which becomes
stronger by decreasing the temperature, belong to this class of materials.
A proof of the role of the magnitude of the applied field on the magnetic
response of the system is shown in Fig. 9.289. 28 for Nd50 Fe40 Al 10lo amorphous
ribbons 25 and 120 IJm in thickness.
120 . ]40
140 r - - - - - - - - - - - - - - - , S5
120 9 T Ribbon 25
RIbbon 2SIlm
11m 5S
100~ ~
?T---.

=3,
. 120 ~
~ 4
~.a.
;80~~4E ~100 ~A -=--
--

u ~
OJ)

5T :f:
/f. 80 3E

$~60u
=3~
~
E60
~ 40 3 T
~403T
IT Y
U
!$
E
60 AST
.a.5T
x
E
2:J::Y
e .3 T
e3T 2:t:
2:J:: ~
~ 40 ~7T
AST
.a.5T +9T
20 1T

o
1T

50
SO 100 150
ISO 200

~ 7T
+9T
250
2S0 300
20 --1.!.
IT
-------
50 100 150 200 250 300

o0 '-----::':SO~-:-l0-:'-:0~-:-lS':-:0----'2~0-=-0-----=2...LSO-=--3::-:'000
0

Temperature (K) Temperature (K)

Figure 9. 28 The magnetization and the maximum of the coercive field vs. temperature
and applied field for Ndso
50 Fe40 Alto
Al lo amorphous melt-spun ribbons (the coercive field is
represented by symbols) .

The larger the applied


appl ied field, the lower the temperature at which the
coercive field attains its maximum. One observes also that the magnetization
exhibits different behavior for applied fields not large enough to align the
magnetic moments parallel to its direction, i. e., to be large enough to
overload the anisotropy energy given by the Nd3+ ions anisotropy.
Thermomagnetic measurements (after cooling without field or after
cooling in an external field) at temperatures above 300 K K indicate the co-
existence of two types of magnetic order: spin-glass-like or cluster-glass-like
short-range order within the clusters and long-range order ferromagnetic order
like a response of the exchange coupling between clusters.
The pair correlation function obtained by Fourier inversion of the
diffraction patterns shows first neighbor peaks at 2. 54, 2. 85 and 3. 36. The
peak at 2. 54 A is at the position expected for the nearest neighbor Fe atoms in
2.54
the dense random packing (DRP)(ORP) model; the other peaks cannot be correlated
to combinations of the radii of Nd (1.82 A), AI (1.43 A) or Fe (1.27 A)
atoms as predicted by the DRPORP model. This disagreement could be ascribed to
the development of a new type of disordered structure.
structure .. Thus, one suggests
that the structure consists in a dense random packing of nanometer-sized
atomic clusters. Consequently, the magnetic structure will consist in non-
collinear short-range magnetic ordered regions randomly distributed. In order to
prove the existence of these short-range magnetic ordered regions (magnetic
clusters) we simulated the topological structure of Ndgo - x Fe x Al 1o - y Siy
amorphous alloys in the shape of ribbons and rods using the reverse Monte
Carlo (RMC) method. The results are shown in Fig. 9.29 9. 29 for different
348 H. Chiriac and N. Lupu

Ndso Fe40 Al lo amorphous samples and in Fig. 9.


Nd50Fe4oAI,o 30 for amorphous and partially
9.30
amorphous ribbons with the nominal compositions Nd55
ss Fe35
Fe3S AI
Al IO
lo and Nd40 Feso
Fe50 AI
Al IO
lo ..

(a) (b)

(e)
(c)

Figure 9. 29 Reverse Monte Carlo (RMC) modeled structures for Nd50 so Fe,o
Fe40 Al,o
AI"
amorphous samples: (a) ribbon 25 ~m; (b) ribbon 120 ~m; (c) rod 1 mm"'
mma> (Nd
atoms are represented by light gray spheres, Fe atoms by black spheres and AI
atoms by dark gray spheres) .

(a) (b)

(e)
(c) (d)

Figure 9. 30 Reverse Monte Carlo (RMC) modeled structures for: Nd 55 ss Fe3s


Fe35 Al,o
AI"
amorphous ribbon 25 ~m (a) and 120 ~m (b): Nd4oFe50AI" amorphous ribbon 25 ~m (c)
(b); Nd'oFesoAl,o
115 ~m (d) (Nd atoms are represented by Iight
and partially amorphous ribbon 115 light gray
spheres, Fe atoms by black spheres and AI atoms by dark gray spheres) .
Bulk Amorphous Magnetic Materials 349

Small Fe-based clusters embedded in a homogenous Nd-based matrix are


observed in amorphous ribbons and cast rods, regarding the composition,
whereas in the partially vitrified ribbon slabs containing Fe and AI and between
them isotropically dispersed Nd and Fe atoms were found, in agreement with
crystall ine compounds (Le Breton et al. ,
the structure proposed for Nd-Fe-AI crystalline
1995).
1995) .

9.66
9. Future Perspectives

Despite the considerable progress achieved in recent years concerning the


knowledge of the atomic structure of amorphous alloys and concerning the
understanding of their basic magnetic properties, several unsolved problems
were encountered in the field of bulk magnetic amorphous alloys. There are
still many questions related to the microstructure of these materials and its
interplay with magnetic properties, especially for Nd-Fe-based high coercivity
bulk amorphous alloys. The magnetic ground states of Nd-Fe-based cluster
amorphous alloys and non-collinear structures existent in these materials are
far from being fully characterized. These problems and others make the study
of the magnetic properties of bulk amorphous alloys a fascinating field of
research.
Bulk magnetic amorphous alloys are interesting also for the field of
engineering applications. For example, the applications of soft magnetic bulk
amorphous alloys are based on the ease of preparation in different bulk shapes
as well as their good soft magnetic properties, comparable with other
amorphous alloys limited to the shape of ribbons, wires, thin films or
powders. They could be used in the future as magnetic cores (they can be
prepared like rings), magnetic heads (the rectangularity of the hysteresis
loop is changing by appropriate thermomagnetic treatments), or like
magnetostrictive materials for sensors.
Although the high-coercivity Nd-Fe-based bulk amorphous alloys are
currently more suitable for fundamental research being below those considered
necessary for economic viability, they could be used successfully in the future
for different appl ications as magnetic recording media, magnetostrictive
applications
materials or low temperature permanent magnets.

References
Allemand, J. , A. Letant, J. M. Moreau, J. P. Nozieres and R. Perrier de la
BiHhie. J. Less-Common Met. 166: 73( 1990)
Bathie.
Alperin, H.A., W.R.
H. A. , W. R. Gillmor, S.J.
S. J. Pickart and J.J.
J. J. Rhyne. J. Appl. Phys.
50: 1958( 1979)
SO:
350 H. Chiriac and N. Lupu

Bozorth, R. M. Ferromagnetism, 3rd edn. IEEE Press. Piscataway, NJ, p.


Bozarth,
444 (1993)
Cahn, R. W. In: H. H. Liebermann ed. Rapidly Solidified Alloys: Processes,
Structures, Properties, Applications. Marcel Dekker, Inc. , New York, p.
1 (1993)
Chen, H.H.S.
S. Acta Metall. 22: 1505
1505(1974)
( 1974)
Chiriac, H., N. Lupu. J. Non-Cryst. Solids 250-252: 751(1999a)
Chiriac, H. and N. Lupu. J. Magn. Magn. Mater. 196-197: 235(1999b)
Chiriac, H. and N. Lupu. J. Magn. Magn. Mater. 215-216: 394(2000a)
Chiriac, H. and N. Lupu. Physica B: Phys. Condens. Matter 299: 293
(2001a)
Chiriac, H. , N. Lupu, R. E. Vandenberghe and K. V. Rao. In: A Inoue, R.
Yavari, W. L. Johnson, R. H. Dauskardt, eds. MRS Proceedings Vol.
644, Supercooled Liquid, Bulk Glassy and Nanocrystalline States of
Alloys. MRS Society, L8.3.1 (2001b)
Chiriac, H.H.,, N. Lupu, F. Vinai, A. Stantero, M. Coisson and E. Ferrara.
Mater. Sci. Forum 360-362: 571(2001c)
Chiriac, H., N. Lupu, K. V. Rao and R. E. Vandenberghe. IEEE Trans.
Magn. 37: 2509(2001d)
Chiriac, H. , N. Lupu, F. Vinai, E. Ferrara, A. Stantero. J. Magn. Magn.
Mater. 226-230: 1379(2001e)
Chiriac, H. and N. Lupu. J. Non-Cryst. Solids 287: 135(20010
Coey, J.M.D.
J. M. D. J. Appl.
App!. Phys. 49: 1646(1978)
1646( 1978)
App!. Phys. Lett. 39: 357 ( 1981>
Croat, J. J. Appl. 1981)
Croat, J. J. IEEE Trans. Magn. MAG-18: 1442( 1982)
Cullity, B. D. Introduction to Magnetic Materials. Addison-Wesley Publishing
Company, p. 385 (1972)
Drehman, A.J., A.L. Greer, D. Turnbull. Appl. Phys. Lett. 41: 716(1982)
Fan, G.J.,
G. J. , W. Loser, S. Roth, J. Eckert and L. Schultz. Appl.
App!. Phys. Lett.
75: 2984( 1999)
Fisher, K. H., J. A. Hertz. Spin Glasses. Cambridge University Press,
Cambridge (1993)
Givord, D. , J. P. Nozieres, J. L. Sanchez L1azamares and F. Leccabue. J.
Magn. Magn. Mater. 111: Ill: 164(1992)
164( 1992)
Hagiwara, M. M.,, A. Inoue and T. Masumoto. Met. Trans. A 13. 373( 1982)
373(1982)
He, Y., S.J. Poon, G.J. Shiflet. Science 241: 1640(1988)
He, Y.,
Y. , C.E.
C. E. Price, S.J.
S. J. PoonandG.J.
Poon and G. J. Shiflet. Phyl. Mag. Lett. 70: 371
(1994)
Hu, B. , J. M. D. Coey, H. Klesnar and P. Rogl. Rogi. J. Magn. Magn. Mater.
117: 225(1992)
225( 1992)
Inoue, A.,
A. , K. Ohtera, A. P.P. Tsai, T.
T. Masumoto. Jap. J. Appl.
App!. Phys. 27:
L479( 1988)
L479(1988)
Inoue, A.A.,, T.
T. Zhang and T. Masumoto. J. J. Non-Cryst. Solids 156-158: 473
473
(1993)
(1993)
Bulk Amorphous Magnetic Materials 351

Inoue, A. Mater. Trans. JIM JI M 36: 866(1995)


866 ( 1995)
Inoue, A., J.S. Gook. Mater. Trans. JIM 36: 1180(1995a)
Inoue, A., J.S. Gook. Mater. Trans. JIM 36: 1282 (1995b)
Inoue, A., Y. Shinohara and J. S. Gook. Mater. Trans. JIM 36: 1427
( 1995c)
Inoue, A. , T. Zhang, W. Zhang, A. Takeuchi. Mater. Trans. JIM 37: 99
( 1996a)
Inoue, A., T. Zhang, A. Takeuchi, W. Zhang. Mater. Trans. JIM 37: 627
(1996b)
Inoue, A., T. Zhang, T. Itoi, A. Takeuchi. Mater. Trans. JIM 38: 359
(1997a)
Inoue, A., T. Zhang and A. Takeuchi. Sci. Rep. RITU. A 44: 261C1997b)
Inoue, A. Bulk Amorphous Alloys: Preparation and Fundamental
Characteristics, Materials Science Foundation, Vol. 4. Trans Tech
Publ ications Ltd. , Switzerland (1998)
Publications
Inoue, A., H. Koshiba, T. Zhang, A. Makino. J. Appl. Phys. 83: 1967
((1998a)
1998a)
Inoue, A., T. Zhang, H. Koshiba, A. Mak Makino.
ino. J. Appl. Phys. 83: 6326
(1998b)
Inoue, A., H. Koshiba, T. Itoi, A. Makino. Appl. Phys. Lett. 73: 744
(1998c)
( 1998c)
Inoue, A., T. Itoi, H. Koshiba, A. Makino. IEEE Trans. Magn. 35: 3355
(1999)
Inoue, A. and H. Kimura. NanoStructured Materials 11. 221 (1999)
Klement Jr., W., R.H. Willens, P. Duwez. Nature 187 869(1960)
Knoch, K. G. , G. Schneider, J. Fidler, E. Th. Henig, H. Kronm lIer. IEEE
Trans. Magn. 25: 3426 ( 1989)
Koshiba, H., H. , A. Inoue, A. Makino. J. Appl. Phys. 85: 5136( 5136(1999)
1999)
Kui, H.W., A.L. Greer, D. Turnbull. Appl. Phys. Lett. 45: 615(1984)
Kui, H.W., D. Turnbull. Appl. Phys. Lett. 47: 796(1985)
Le Breton, J. M. , J. Teillet, D. Lemarchand, V. De Pauw. J. Alloys Compo
218: 31(1995)
31C 1995)
Lupu, N. and H. Chiriaco Mater. Trans. 42: 670(2001a)
Lupu, N. , H. Chiriac, A. Takeuchi and A. Inoue. In: Sara A. Majetich, John
Q. Xiao, Manuel Vasquez, ed. MRS Proceedings Vol. 674. Application of
Ferromagnetic and Optical Materials, Storage and Magnetoelectronics,
MRS Society, U.2.7.1.(
U. 2. 7.1. ( 2001b)
Lupu, N. , R. Delaplane, R. L. McGreevy and H, Chiriac. Chiriaco Appl. Phys. A-
Mater. 74: 5680 (2002)
Ma, L. , A. Inoue. Mater. Lett. 38: 58( 58 ( 1999)
Masumoto, T., A. Inoue and M. Hagiwara. U. S. Patent No.4, 523, 626
(1985)
Matsubara, E., E. , T. Zhang and A. Inoue. Sci. Rep. RITU A43, 83( 1997)
83(1997)
McGreevy R. L. and L. Pusztai. Mol. Simulation 1: 359 (1988) ( 1988) (more
352 H. Chiriac and N. Lupu

information available on http://www.studsvik.uu.se)


R. K. J. Magn. Magn. Mater. 54-57: 450(1986)
Mishra, R.K. 450( 1986)
Moorjani, K. and J. M. D. Coey. Magnetic Glasses. Elsevier, Amsterdam
(1984)
0' Connor, A. S., L. H. Lewis, R. W. McCallum, K. W. Dennis, M. J.
Kramer, D. T. Kim Anh, N. H. Dan, N. H. Luong and N. X. X Phuc. In: H.
Kaneko, M. Homma, M. Okada, eds. Proceedings of the Sixteenth
International Workshop on Rare-Earth Magnets and Their Applications. The
p. 475 (2000)
Japan Institute of Metals, Sendai, Japan, p.475
Ortega-Hertogs, R. J. Ph. D. Thesis. Royal Institute of Technology.
Stockholm, Sweden (2000)
Peker, A. , W. L. Johnson 2342 ( 1993)
Johnson. Appl. Phys. Lett. 63: 2342(
Oi, 0., H. Sun, R. Skomski and JJ. M. D D. Coey. Phys. Rev. B 45: 12,278
(1992)
( 1992)
Schwarz, R B. In: H H. H. Liebermann, ed. Rapidly Solidified Alloys:
Processes, Structures, Properties, Applications. Marcel Dekker, Inc.,
New York, p. 157 (1993)
Shingu, P. H. and K. N. Ishihara. In: H. H. Liebermann, ed. Rapidly
Solidified Alloys: Processes, Structures, Properties, Applications. Marcel
Inc. , New York, p. 103 (1993)
Dekker, Inc.,
Wang, X.Z., Y. Li, J. Ding, L. Si, H.Z. Kong. J. Alloys
AlloysComp.
Compo 290: 209
(1999)
Wannberg, A. , A. Mellergard, P. Zetterstram, R. Delaplane, M. Granros,
Gramos,
L. -E. Karlsson and R. L. McGreevy. J. Neutron Research 8: 133 ( 1999)
Wiesner, H. and J. Schneider.
Schneider Phys. Status Solid (a) 26: 71(1974)
Yoshizawa, Y., S. Oguma and K. Yamauchi. J. Appl. Phys. 64: 6044
(1988)
Index

alignment 49,51,
49, 51, 185,289,293,294,
185, 289, 293, 294, glass-forming ability 304,305,311,313
333,343 - 315,319,325
315 ,319,325 - 327,329,344
327 ,329,344
amorphous alloys 97,101,
97, 101, 115, 118, grain 3,5,8,9,14,18,25,69,74 -76,
121,128,132,134 - 136,138,139,148, 78,85,
78,85,88,92,
88, 92, 100, 102 - 107, 109, 111, 111 ,
150,160,165,234,303-306,309,311-
150, 160, 165,234,303 - 306,309,311 - 117, 118, 128 - 132,134,140,152,153,
117,118,128 132, 134, 140, 152, 153,
315,318 - 321 ,323 - 325,327,329,332
325 ,327,329,332 - 204,249
336,340,341,343-347,349
336,340,341,343 - 347 ,349 Hall effect 277 ,295,296
anisotropy 1 ,4,8, 11
1,4,8, 11 - 14, 19 - 21 ,24, HDDR 1- 3 , 5 - 7 , 9 - 11,
11 , 13 - 16, 18,
25,78, 102, 103, 110, 115 - 118, 120 - 19,21 -25,39,50,67
19,21-25,39,50,67
122,129,130,133 - 145,150,151,156,
145, 150, 151,156, Heusler alloys 267 - 27
2711
159 - 161,163
161 , 163 - 165,167,168,194,195,
165, 167, 168, 194, 195, hot compression 49,51
49 ,51 ,52
197,198,200,201,211,217,228,245 - hydrogen absorption 4,5,21 ,24
251,254,257,258,260,261,276,277,287 isotropic 2,22 - 25,34 - 37,40,47,48,
-289,291,312,323,324,336,338,343,
- 289 , 291 ,312,323,324,336,338,343, 50,52,55,57,58,62-64,106-108,110,
50,52,55,57,58,62 - 64 , 106 - 108,110,
345- 347 135,182,247,337,349
bonded magnets 2,3,23,25,32 - 34 34,37
,37 laser 69 - 75 ,77 ,78,80,82 - 85, 85,183
183
-41,44-
- 41 ,44 - 47,49,51,53,55,57,60
47,49,51 ,53,55,57,60 - 63, lithography 69 -71,80,85
- 7 1,80 ,85
106,107 magnetic bistability 119,120,122,123,
core-shell microstructure 237 125,166-169
corrosion 44,46,47,57,60,61,107,
44,46,47,57,60,61, 107, magnetic easy axis 78, 116
110,115,303
110 , 115 ,303 magnetic excitations 291 ,293
crystall ization 89,90,92,93,96,97, 100 magnetic permeability 118,119,142,
118, 119, 142,
-106,111,117,118,122,128,130,131,
- 106, 111, 117, 118, 122, 128, 130, 131 , 143,154,155,159
143, 154, 155, 159 - 162,164,165,271,
162, 164, 165, 271 ,
133, 134, 139 - 141,151,152,222,275,
133,134,139 141 , 151 , 152, 222, 275, 305,312,313,319-321,345
278,303,305,306,311,312,315,317
278,303, 305, 306, 311 , 312, 315, 317 - magnetic sensors 118,119,166 - 168,
319,321 ,325 - 328 ,333,334,341 ,345 172
disproportionation 3,5 - 8, 10, 12 - 16, magnetic transitions 346
18,19,21,22,24,25,39,222,223,226, magnetization curve 36 - 40, 42, 55,
235 108,120,121, 135, 193, 195, 197,201,
108,120,121,135,193,195,197,201,
dynamic magnetic properties 218,234, 286,335,337,340,341
241,252 magnetostriction constant 119, 120, 122
electrical resistivity 6, 194,293,312,
194, 293, 312, -125,128,134,142,145,147,149,152-
- 125,128,134,142,145,147,149,152-
317,318 154,161,164-
154 , 161 , 164 - 166,312,314,320
166 ,312 ,314 ,320
extrusion 32,34,36,45,53,55,56,64 Mbssbauer spectra
Mossbauer 312,342,343
ferromagnetic transition 305 nanocomposite 38, 88 - 90, 92, 93,97,
giant magneto-impedance 119,
119 , 166 99-
99-111,115
111,115
354 Index

nanocrystalline materials 116, 118, 119, polyol process 218, 222 - 224, 226,
133,134,153,165,166
133, 134, 153,165, 166 228,230,232,242,248,260,261
Nd, Fe" B
Nd,Fe14B 2-6,8-10,12-15,16,20
2 - 6,8 - 10, 12 - 15, 16,20 PtMnSb
PtMn8b 267,269,271 ,272,274
,272 ,274 - 277
- 21 ,23,25,38,88 - 91 ,93,96 - 98, 100 rapid quenching 115, 120, 123,221 ,324
-110 8mCo
SmCo 2,34,40-42,45,47,51-53,
2,34,40 - 42,45,47,51 - 53,
Nd3 (Fe, Tj)29 2,24 62,63
neutron diffraction 6,272,298,306,308 8m, Fe" N,
Sm,Fe17N, 2,25
- 310,329,332
310,329 ,332 stability 44,57,60,63,75,109,110,
NiMnSb
NiMn8b 267 - 287 ,289 - 292,294,295, 182,213,303,311,313,314,316,319,
182, 213, 303, 311 , 313, 314, 316, 319,
298 326,341
nucleation 12,20,75,91, 120, 125 - static magnetic properties 218, 242,
128,207, 209, 219 - 224, 230, 231 , 242, 245,247
261,282,285,321 suction casting technique 305
permanent magnet 1, 2, 4, 23, 24, 40, thin films 74,78,85, 110, 121 ,268,275,
74,78,85,110,121,268,275,
67,78,88 - 90,92,97,99, 104 - 111 ,185,
67,78,88-90,92,97,99,104-111,185, 277 , 279, 280, 287, 292, 298, 303, 306,
324,349 313,349
phase transformation 69, 74 - 78,98,
69,74 78, 98, ThMn12-type
ThMnl2 -type 2,23,24
151 tunnel magnetoresistance 184,267,278
Handbook of Advanced
Magnetic Materials
Volume IV: Advanced Magnetic Materials:
Properties and Applications
Handbook of Advanced
Magnetic Materials
Volume IV.- Advanced Magnetic Materials:
Properties and Applications

Edited by:

Yi Liu
Center for Materials Research and Analysis
University of Nebraska
Lincoln, Nebraska

David J. Sellmyer
Center for Materialsl Research and Analysis
University of Nebraska
Lincoln, Nebraska

Daisuke Shindo
Institute of Multidisciplinary Research for Advanced Materials
Tohoku University
Sendai, Japan

(A) Tsinghua University Press i2J Springer


Library of Congress Cataloging-in-Publication Data

ISBN-IO: 1-4020-7983-4 e-ISBN-IO: 1-4020-7984-2


ISBN-I3: 978-1402-07983-2 e-ISBN-13: 978-1402-07984-9

2006 Springer Science+Business Media, Inc.


All rights reserved. This work may not be translated or copied in whole or in part without the
written permission of the publisher (Springer Science+Business Media, Inc., 233 Spring Street, New
York, NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis.
Use in connection with any form of information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter developed is
forbidden.
The use in this publication of trade names, trademarks, service marks and similar terms, even if they
are not identified as such, is not to be taken as an expression of opinion as to whether or not they are
subject to proprietary rights.

Printed in the United States of America.

9 8 7 6 5 432 I SPIN 11097730

springeronline.com
Handbook of Advanced Magnetic Materials
Preface

In December 2002, the world's first commercial magnetic levitation supertrain


went into operation in Shanghai. The train is held just above the rails by
magnetic levitation (maglev) and can travel at a speed of 400 km/hr
completing the 30km journey from the city to the airport in minutes. Now
consumers are enjoying 50 GB hard drives compared to 0.5 GB hard drives ten
years ago. Achievements in magnetic materials research have made dreams of
a few decades ago reality. The objective of this book is to provide a
comprehensive review of recent progress in magnetic materials research. The
whole book consists of four volumes, each volume focusing on a specific field.
Graduate students and professional researchers are targeted as the readers.
Each chapter will have an introduction to give a clear definition of basic and
important concepts of the topic. The details of the topic are then elucidated
theoretically and experimentally. New ideas for further advancement are then
discussed. Sufficient references are also included for those who wish to read
the original work. Many of the authors are well known senior scientists. We
have also chosen some accomplished young scientists to provide reviews on
new and active topics.
In the last decade, one of the most significant thrust areas of materials
research has been nano~tructured magnetic materials. There are several
critical sizes that control the behavior of a magnetic material. For example,
the coercivity of a magnetic material made of particles increases with
decreasing particle size, reaching a maximum where coherent rotation of a
single-domain particle is realized, and then decreases with further decrease of
the particle size. For a composite made of a magnetically hard phase and soft
phase, when the grain size of the soft phase is sufficiently large, the soft and
hard phases reverse independently. However, when the grain size of the soft
phase is reduced to a size of about twice the domain wall thickness of the hard
VI Preface

phase, the soft and hard phases will be exchange-coupled and behave as if a
single magnetic phase is present. Such behavior can be used to increase the
energy product of high-performance permanent magnets. Size effects become
critical when dimensions approach a few nanometers, where quantum
phenomena appear. The first volume of the book has therefore been devoted
to the recent development of nanostructured magnetic materials, emphasizing
size effects.
Our understanding of magnetism has advanced with the establishment of
the theory of atomic magnetic moments and itinerant magnetism. In general,
the magnetism of a bulk material can be considered as the superposition of
atomic magnetic moments plus itinerant magnetism due to conduction
electrons. In practical applications the situation becomes much more
complicated. The boundary conditions have to be taken into account. This
includes the size of the crystals, second-phase effects and intrinsic properties
of each phase. The effects of magnetic relaxation over long periods of time
can be critical to understanding. Simulation is a powerful tool for exploration
and explanation of properties of various magnetic materials. Simulation also
provides insight for further development of new materials. Naturally, before
any simulation can be started, a model must be constructed. This requires that
the material be well characterized. Therefore the second volume of the book
provides a comprehensive review of both experimental methods and simulation
techniques for the characterization of magnetic materials. After an introdJction, each
section gives a detailed description of the method and the following sections
provide examples and results of the method. Finally further development of the
method will be discussed.
The success of each type of magnetic material depends on its properties
and cost which are directly related to its fabrication process. Processing of a
material can be critical for development of artificial materials such as
multilayer films, clusters, etc. Moreover, cost-effective processing usually
determines whether a material can be commercialized. In recent years
processing of materials has continuously evolved from improvement of
traditional methods to more sophisticated and novel methods. The objective of
the third volume of the book is to provide a comprehensive review of recent
developments in processing of advanced magnetic materials. Each chapter will
have an introduction and a section to provide a detailed description of the
processing method. The following sections give detailed descriptions of the
processing, properties and applications of the relevant materials. Finally the
potential and limitation of the processing method will be discussed.
The properties of a magnetic material can be characterized by intrinsic
Preface VB

properties such as anisotropy, saturation magnetization and extrinsic


properties such as coercivity. The properties of a magnetic material can be
affected by its chemical composition and processing route. With the continuous
search for new materials and invention of new processing routes, magnetic
properties of materials cover a wide spectrum of soft magnetic materials, hard
magnetic materials, recording materials, sensor materials and others. The
objective of the fourth volume of this book is to provide a comprehensive
review of recent development of various magnetic materials and their
applications. Each chapter will have an introduction of the materials and the
principals of their applications. The following sections give a detailed
description of the processing, properties and applications. Finally the
potential and limitation of the materials will be discussed.
NASA is considering the launching of spacecraft by maglev. The first
stage rocket, which accounts for two-thirds of the cost and is lost every
launch, would be replaced by a maglev track. Using a 50 ft track NASA
scientists have accelerated a model spacecraft to 96kph in less than half a
second. In the last few decades the knowledge of mankind has been expanding
rapidly into deep space measured by light years and the nano world where
building blocks of atoms are being engineered. Magnetism and magnetic
materials are among the most intriguing and fascinating science and
engineering fields. Undoubtedly advances in magnetic materials research will
continue to fuel our understanding of the universe in the new century. We hope
this book will provide a useful reference for researchers working at the frontier
of magnetic materials research.
We would like to express our sincere thanks to all our devoted authors,
technical editors, and publishers for making this book possible.

The editors
Contents

Preface V
List of Contributors :x.vn
1 Recent Developments in High-Temperature Permanent Magnet
Materials .
1. 1 Introduction '" .
1.1.1 Requirement for High-Temperature Permanent
Magnet Materials , .
1. 1.2 Review of Conventional Permanent Magnet
Materials .
1.2 Candidate Alloy System for High-Temperature Permanent
Magnet Materials 4
1.3 New Sm2Co17-Based High-Temperature Permanent Magnet
Materials ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 7
1.3. 1 Effects of Compositions on High-Temperature Intrinsic
Coercivity of Sm2 -TM 17 Permanent Magnets ... ... ... ... ... 7
1. 3. 2 Magnetic Properties of New Sm2 C0 17 -Based
High-Temperature Permanent Magnet Materials 13
1.3.3 Dynamic Characterizations of New High-Temperature
Permanent Magnets 16
1.3.4 Long-Term Thermal Stability of New High-Temperature
Permanent Magnets ... ... ... ... ... ... ... 19
1 .3. 5 Microstructure and Crystal Structure 21
1. 3.6 Coercivity in Sm2 -TM 17 -Type Permanent Magnets 23
1. 4 New Approach to Calculate Temperature Coefficients of
Magnetic Properties ... ... ... ... ... 35
1. 5 SmC07-Type Permanent Magnets 38
1.6 Temperature Compensated Rare Earth Permanent Magnets and
Modeling of Temperature Coefficient of Magnetization ... ... ... ... 39
1.7 Other High-Temperature Permanent Magnet Materials
and Furture Perspective 46
References 47

2 New Rare-Earth Transition-Metal Intermetallic Compounds and


Metastable Phases for Permanent Magnetic Materials 50
2. 1 Introduction'" ... ... ... ... ... ... ... ... ... ... ... ... ... ... 50
X Contents

2. 1. 1 Overview of Structures ... 51


2.2 Experimental Processing '" 54
2.2.1 Melting 54
2.2.2 Mechanical Alloying/Milling 55
2.2.3 Rapid Quenching 55
2. 2. 4 HDDR ... ... ... ... 55
2.2.5 Sol id-State Reaction 56
2. 2. 6 Solid-Gas Reaction 56
2.2.7 Sputtering .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. 56
2.3 New Intermetallic Compounds 57
2.3.1 RCFe, T ,X)'2CX=B,C) 57
2.3.2 R2Fe14BN6 59
2.3.3 R6Fe13-xMl+x 61
2.3.4 R3 T29 Si 4 B,o 61
2.4 Metastable Phases 64
2. 4. 1 RFes ... ... 64
2.4.2 RT 7 CT=Fe,Co,Ti,etc.) 65
2.4.3 RT 7 N6 68
2.4.4 R-Fe-C 72
2.4.5 Nd2Fe14BN6 74
2.4.6 RFe12 75
2.5 Discussion 76
2. 6 Conclusions'" 79
References '" 80

3 Magnetic Properties and Interstitial Atom Effects in the RCFe,M)12


Compounds ... ... ... 86
3. 1 Introduction .. 86
3.2 Interstitial Atom Effects and Intermetallic Compounds 87
3.2. 1 Crystallographic Structures 87
3.2.2 Curie Temperature and Saturation Magnetization 93
3.2.3 Magnetocrystall ine Anisotropy 95
3.2.4 Hyperfine Interactions ...... 97
3.3 Origin of the Interstitial Atom Effects 101
3.3. 1 Crystal Field Effect 101
3. 3.2 Band Structures 103
3.4 Novel Permanent Magnets Based on 1 : 12 Nitrides 108
3.4. 1 The Domain Structures in Compounds, Nitrides
and Carbides 108
3. 4.2 Hard Magnetic Properties of Nitrides 111
3. 5 Conclusions and Prospect .. .. .. 118
References ... ... ... ... ... ... ... ... ... ... 119
Contents XI

4 Nanocrystalline Soft Magnetic Materials and Their Applications 124


4. 1 Introduction ... ... ... ... ... ... ... ... ... ... ... 124
4.2 Magnetic Properties and Microstructure 125
4. 2. 1 Magnetic Properties 125
4.2.2 Microstructure .. ... ... 133
4. 2. 3 Origin of Magnetic Softness 136
4. 3 Applications ... ... ... ... ... ... ... ... ... 143
4. 3. 1 Parts for Noise Measures 145
4.3.2 Magnetic Parts for Communications 147
4. 3 . 3 Parts for Pulsed Power ... 148
4.3.4 Parts for Power Supplies 149
4. 3. 5 Current Sensors ... ... ... ... 152
4. 3.6 Electromagnetic Shielding, Wave Absorbers 152
4.4 Conclusions 155
References ... ... ... 155

5 Spin-Density Waves and Charge-Density Waves in Cr Alloys 159


5. 1 Introduction ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 159
5.2 Magnetic Properties of Antiferromagnetic Cr and Cr Alloys 161
5.2. 1 Magnetic Phase Diagrams 161
5. 2.2 Magnetic Phase Transition 164
5. 3 Local Moments in Cr Alloys 168
5.3.1 Magnetic Susceptibility 168
5. 3. 2 Local Moments in Non-Magnetic Impurities-Local
Spin-Density Waves ... ... ... ... ... ... ... ... ... ... ... ... ... ... 170
5. 3. 3 Correlation with Giant Magnetoresistance and Exchange
Bias in Cr Alloys 174
5.4 Spin-Glass Phase in Cr-Alloys 177
5 . 5 Neutrons and X-Ray Scattering: CDW/ SW and SDW Domains 180
5.5.1 Technical Antiferromagnetism 180
5.5.2 Neutrons and X-Ray Two Complementary Techniques
Used in Studies of Cr Magnetism ...... ... ... ...... ... ...... 182
5.5.3 Aspects of Neutron and X-Ray Scattering in Cr and
Cr-Alloys 183
5.5.4 Orbital Moment in Cr and Resonant Scattering 185
5.5.5 Magnetic Q Domains in Cr COOD Face 193
5. 5.6 Paramagnetic Critical Fluctuations in Cr and
in Cr-0.2at. %V 195
5. 6 Ultrasonic Measurements in Cr and Cr-Alloys 197
5. 7 Bulk Chromium Properties Relevant for Thin Film Devices 202
5. 8 Final Remarks 204
References ... ... ... ... 205
XII Contents

6 New Magnetic Recording Media 211


6. 1 Introduction ... ... ... ... ... ... 211
6.2 New Media Development by Sputtering 213
6.2. 1 Introduction'" ... ... ... 213
6. 2 . 2 Hexagonal Co-Alloys 215
6 . 2 . 3 L 10 Alloys 216
6.2. 4 Rare Earth-Transition Metal Compounds 216
6.2.5 Nanocomposite Films 217
6 . 2 . 6 AFC Media 218
6. 3 Chemical Synthesis 219
6.3.1 Introduction 219
6.3.2 Stabilization of Nanoparticle Dispersion 220
6.3.3 Chemical Synthesis of Monodisperse
Nanoparticles ... ... ... ... ... ... ... ... 221
6. 3.4Magnetic Nanoparticle Assembly 225
6. 3.5Conclusions 227
6. 4 Self-Assembly ... ... 228
6. 4. 1 Introduction'" 228
6.4.2 Porous Alumina Template Technique 229
6. 4.3 Self-Assembled Nanowire/Dot Arrays as Recording
Media-Merits and Challenges .. .. .. .. 235
6. 5 Concluding Remarks 235
References ... ... ... ... ... ... 236

7 Magneto-Optical Properties of Nanostructured Media .. 241


7.1 Introduction 241
7.2 Electromagnetic Theory of the Magneto-Optical Effect 242
7.2. 1 Faraday Rotation .. .. 242
7.2.2 Kerr Rotation 244
7.3 Microscopic Models 247
7.3.1 Interband Transitions 247
7. 3.2 Intraband Transitions 248
7 . 4 Magneto-Optical Measurements 250
7. 4. 1 Experimental Principles 250
7.4.2 Description of the Experimental System 253
7. 5 Magneto-Optical Spectra of Magnetic and Non-Magnetic
Metals 255
7. 5. 1
The Magneto-Optical Effect in the Noble Metal Ag 255
7.5.2Magnetic Granular Films 256
7. 5.3Magnetic Co-Cu Multilayers 263
7.6 Summary '" ." 268
References ... ... ... ... 268
Contents :xm

8 Magnetoresistive Recording Heads 271


8. 1 Introduction ... ... ... ... ... ... 271
8. 2 Physics of Magnetoresistance 274
8.3 Magnetoresistive Head Fabrication 280
8. 4 Noise in Magnetoresistive Sensors 280
8. 5 Future of Magnetic Data Storage: Challenge in Maintaining
Rapid Pace of Development 283
Appendix 8. 1 Glossary of Commonly Used Industry Terminology'" 284
References ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... 288

9 Magnetic Random Access Memories for Computer Data Storage 292


9. 1 Introduction ... ... ... ... ... ... ... ... 292
9.2 MRAM Using Spin-Valve Effect 293
9. 3 EPROM Using Hall Effect 296
9.4 MRAM Using Spin-Dependent Tunneling Effect 297
9. 5 Future Perspective 304
References ... ... ... ... ... ... 305

10 Magnetoresistive Thin Film Materials and Their Device


Applications ... ... ... 307
10. 1 Introduction 307
10.2 Magnetoresistive Materials 308
10.2.1 AMR Materials 310
10.2.2 Giant Magnetoresistive Materials 312
10.2.3 Tunnel Magnetoresistance 314
10.2.4 Colossal Magnetoresistance 319
10.2.5 Advanced Magnetoresistive Materials 320
10. 3 Magnetoresistive Devices ... ... ... 321
10 . 3. 1 Magnetic Field Sensors 322
10.3.2 Magnetic Field Gradient Sensors'" 324
10. 3. 3 Electrical Current Sensors 324
10. 3.4 Integrated Magnetic Sensors 325
10. 3.5 Galvanic Isolators 326
10. 3.6 Advanced Magnetoresistive Devices 329
10.3.7 Applications of Magnetic Field Sensors 331
References ; 333

11 Nano-Structural Magnetoelastic Materials for Sensor Applications 339


11 . 1 Introduction 339
11.2 Metallic Glass Preparation and Properties 341
11. 2. 1 Preparation of Metallic Glasses 341
11 .2.2 Summary of Physical and Magnetic Properties of
XIV Contents

Metallic Glasses '" '" , 342


11.2.3 Tailoring of Metallic Glasses for Sensor
Applications 346
11.3 Sensor Applications 348
11. 3. 1 Induction Sensors with Metallic Glass Cores 349
11 . 3.2 Giant Magneto-Impedance / Stress-Impedance
Sensors ... ... ... ... ... ... ... ... ... ... ... ... 354
11. 3.3 Magnetostrictive Delay Line Sensors 356
11. 3. 4 Magnetoelastic Resonance Sensors 359
11 . 3.5 Magneto-Surface-Acoustic-Wave Sensors 368
11.3.6 Gyro-Magnetic Sensors with Metallic Glass
Cores 369
11 . 4 Conclusions " '" 371
References 372

12 Soft Magnetic Films and Wires for Magnetic Field Sensors 380
12. 1 Introduction ... ... ... ... ... ... 380
12.2 Linear Magneto-Impedance 383
12.3 Electrodynamic Problems of Linear Magneto-Impedance 392
12. 4 Nonlinear Magnetization Reversal Induced by ac Current 397
12.5 Manufacturing of Magneto-Impedance Units for Magnetic
Field Sensors 404
12.6 Magneto-Impedance Sensors of Magnetic Field 408
12. 7 Some Concluding Remarks 409
References 410

13 Microwave Permeability of Magnetic Films 414


13. 1 Introduction 414
13.2 High Frequency Permeability of a Single Domain Ellipsoid 415
13.2. 1 Landau-Lifshitz Equation and Frequency Dispersion of
Permeability .. .. .. '" 415
13. 2 . 2 Snoek' sLaw 417
13.2.3 Relation Between the Static Permeability and
the Resonance Frequency in Thin Films 418
13.2.4 The Effect of Damping 420
13. 2.5 Integral Relations 421
13.3 Models of Microwave Permeability of Actual Thin Films 424
13. 3. 1 The Effect of Out-of-Plane Anisotropy'" 424
13.3.2 The Effect of Stripe Domain Structure 425
13.3.3 The Effect of Finite Film Thickness 427
13. 3.4 Account for the Eddy Currents 428
13.3.5 Permeability of Patterned and Granular Films 430
Contents XV

13.4 Experimental Data on the Permeability of Thin Films 431


13.4.1 Permeability Measurement Techniques 431
13.4.2 Experimental Microwave Magnetic Spectra 433
13.5 Microwave Appl ications for High Permeable Materials 440
13. 6 Conclusions ... ... ... ... ... 442
References ....................... '" 443

Index 446
List of Contributors

1. Sam Lin University of Dayton


Liu@udri. udayton. edu
2. Z. D. Zhang Shenyang National Laboratory for Materials
Science, Institute of Metal Research and Inter-
national Center for Physics, Chinese Academy
of Science, Shenyang 110016, P. R. China
3. Jinbo Yang University of Missouri-Rolla, Rolla M065409,
USA
yangj@umr.edu
Yingchang Yang
4. Yoshihito Yoshizawa Hitachi Metals, Japan.
yoshihito_yoshizawa@po. hitachi-metals. co. jp
5. A. J. A. de Oliveira Departamento de Fisica-Universidade Federal
de Soo Carlos, CP 676- 13565-905,
Soo Carlos, SP Brazil
adilson@df. ufscar. br
P. C. de Camargo Departamento de Fisica-Universidade Federal
do Parana, CP 19044-81. 531-990, Curitiba-
PR, Brazil
camargo@fisica. ufpr.
6. D. J. Sellmyer et at. UNL. USA
dsellmye@unlnotes. un!. edu
7. Liangyao Chen Fudan University, P. R. China
Iychen@fudan. ac. cn
Songyon Wang Sywang@fudan.ac. cn
Roger D. Kirby Department of Physics University of Nebraska
Lincoln, Nebraska 68588-0111
Rkirby1 @un!. edu
8. Christopher D. Keener Senior Engineering Manager, Hitachi Global
Storage Technologies
Hail iang Storage Products Company, Ltd.
1IF, Block C, Great wall Building Science &
Industry Park Nanshan District Shenzhen 518057
P. R. China
XVDI List of Contributors

7. Hiroaki Kato Department of Applied Physics, Graduate School


of Engineering, Tohoku University, Aoba-yama
08, Aoba-ku, Sendai 980-8579, Japan
Mitsuhiro Motokawa Institute for Materials Research, Tohoku Universi-
ty, 2-1-1 Katahira, Aoba-ku,
Sendai 980-8577, Japan
8. S. Wirth Max-Planck-Institute for Chemical Physics of Sol-
ids, N6thnitzer Str. 40, 01187
Dresden, Germany
S. von Molnar MARTECH, Florida State University, Tallahas-
see, FL 32306-4351, USA

9. Kiyonori Suzuki School of Physics and Materials Engineering,


Monash University, Victoria 3800, Austral ia
1 Recent Developments in High-Temperature
Permanent Magnet Materials

Sam CShiqiang) Liu

1. 1 Introduction

1.1.1 Requirement for High-Thmperature Permanent Magnet Materials


Permanent magnet materials capable of reliably operating at high
temperatures (~400 C) are required for future advanced power systems in a
proposed more electric aircraft (MEA) initiative. A major objective of the MEA
initiative is to increase aircraft reliability, maintainability, and supportability,
including drastically reducing the need for ground support equipment (Fingers
and Rubertus, 2000). This advancement will be accomplished in part through
the development of advanced power components such as magnetic bearings,
integrated power units, and internal starter/generators for main propulsion
engines. New high-temperature magnets are enabling technologies for the
development of these new power components. Power system designers
frequently find that magnetic materials impose technological limitations on their
designs. Compromises are generally required between the desired
performance and the magnetic, mechanical, and electrical properties of
available materials. If new materials can operate at ~ 400 "0, then new
advanced designs will be possible. Air-cooling, rather than complicated liquid
cooling and its necessary logistics support, will become an operational
capability. Likewise, oil-Iess/lubeless gas turbine engines and space power
systems will be possible.

1. 1. 2 Review of Conventional Permanent Magnet Materials


Currently, the most widely-used permanent magnet materials are Alnico, hard
ferrites, and high-performance rare earth-transition metal (RE-TM) permanent
magnets, including SmCos ' 8m2 (Ce, Fe, Cu, Zr) 17 (8m2 TM 17 or 2 : 7), and Nd-
Fe-B magnets. Table 1. 1 lists the magnetic properties of some commercial
magnets at room temperature. Figures 1. 1 and 1. 2 summarize the temperature
2 Sam (Shiqiang) Liu

dependence of the maximum energy product, (BH)max' and intrinsic


coercivity, MH c ' for five types of commercial permanent magnetic materials
(Walmer et al. ,2000a). As shown in these figures, the magnetic properties of
rare earth magnets are superior to all other magnetic materials. Among rare
earth magnets, Nd-Fe-B-type magnets have the highest static (BH ) max at
temperatures T < 130C. At T> 130 'C ,8m2 TM 17 (2 : 17) magnets have the
highest static (BH)max' The only magnet that maintains relatively high intrinsic
coercivity (-10 kOe) at temperatures up to 300 'C is 8m2 TM 17

Thble 1.1 Magnetic properties of some corrmercial permanent magnets at room temperature.
Magnet B,(kG) MHc(kOe) (BH)max (MGOe) TeCC)
Alnico 5 12 0.5 5 840
Hard ferrite 3.5 2.5 3 450
8mCos 9 30 20 727
8m2 TM 17 11 30 28 820
N~Fe14B 14 15 40 312

60

oo---.50
6
~ 40

~tf
.g 30
ec..

"ai 20
E
::>
E
.~ PtCo
~ 10 V-:~-J-~I7----";;,--,,--~
Alnico 5
x:~~~=-?"",,-_:>---o--c~=li1---o
SrFeI20j~19~-~_f--"*-~
O'-_---'_ _--'-_ _--'-_ _.....L.---==----c-'-:-_
-200 -100 0 100 200 300
Temperature CC )
Figure 1. 1 Temperature dependence of (BH) max for five types of permanent magnets.

Possessing the highest Curie temperature and moderately high


magnetization and energy product among the high-performance rare earth
permanent magnets, 8m2 TM 17 magnets are the best conventional high-
temperature permanent magnets (Ray and 8trnat, 1972; Ray, 1984; Ray and Liu,
1992). A conventional 8m2 TM 17 magnet can operate up to 300'C . The problem
Recent Developments in High-Temperature Permanent Magnet Materials 3

30

oL~::j:::~::::::L:::~=:1.:~~~~
-1 SO -SO 0 SO 1SO 250 350
Temperature (C )

Figure 1.2 Temperature dependence of MHo for five types of permanent magnets.

associated with higher-temperature (> 300 "C) operation has been that the
intrinsic coercivity of these magnets drops sharply with increasing
temperature. Upon heating, MHo of the 2 : 17 magnets drops sharply from their
room temperature values of 20 to 30 kOe (or higher) to only 3 to 6 kOe at 400 C
and 1 to 3 kOe at 500C (Fig. 1.3). Low intrinsic coercivity at high temperatures
35

Best conventional high-temperature


permanent magnet (2: \7)
~
Q)

8 25
6
~
E 20
~
g \5
0
S,
~10
80
~ 5

o 100 200 300 400 500


Temperature CC )

Figure 1. 3 Temperature dependence of the best conventional high-temperature 2 : 17 magnet.


4 Sam (Shiqiang) Liu

results in nonlinear 2nd-quadrant induction demagnetization curves (8 curves)


above::::::::: 300 C. A linear 2nd-quadrant 8 curve is critical for all dynamic
applications, such as for generators and motors.
In a dynamic application, the operating point of a magnet keeps cycling. If
the intrinsic coercivity is low, then the induction demagnetization curve can be
nonlinear. Under this circumstance, the operating point of the magnet can be
reduced to below the knee in the induction demagnetization curve and the
induction can be significantly reduced irreversibly. If the intrinsic coercivity of
the magnet is sufficiently high, then the induction demagnetization curve can be
linear. Under this circumstance, the induction will be reversible around the
operating point even at a quite low permeance value as shown in Fig. 1.4. The
maximum operating temperature of a magnet can be defined as the
temperature limit at which the induction demagnetization curve of the magnet
maintains the linearity. Therefore, to increase the operating temperature of
permanent magnet materials, the key is to increase intrinsic coercivity at high
temperature, so that their induction demagnetization curves remain linear at
the operating temperature.

Figure 1. 4 Intrinsic coercivity and linearity of an induction demagnetization curve.

1. 2 Candidate Alloy System for High-Temperature Pennanent


Magnet Materials

Two criteria of high-temperature permanent magnets are high Curie


temperature and high magneto-crystalline anisotropy. At least within the
foreseeable future, these requirements can be met probably only in the Sm-Co-
based system. It is well known that in a RE-TM compound, the Curie
temperature is primarily determined by the TM sublattice , while the crystalline
Recent Developments in High-Temperature Permanent Magnet Materials 5

anisotropy is primarily determined by the RE sublattice, except at


temperatures as high as close to the Curie point. Research on RE-TM
compounds has indicated that among all 3d transition metals, Co provides the
highest Curie temperature, while among all light rare earths, Sm usually
provides the highest crystall ine anisotropy. One exception is the cubic Laves
1 : 2 compounds for which the Fe compounds have higher Curie temperature
than the Co compounds.
Figure 1. 5 shows Curie temperature, T c' versus Co content in Sm-Co
binary compounds. In this figure, Curie temperature data for LaCo 13 and
Sm2 C0 14 B are also included. It can be seen from Fig. 1. 5 that there exists an
almost linear relationship between the Curie temperature and the Co content:
the higher the Co content in a compound, the higher the Curie temperature of
the compound, which clearly demonstrates the importance of the relationship
of the Co content to the Curie temperature. Figure 1. 6 illustrates the
dependence of the anisotropy field, H., on individual rare earth in binary
RECos compounds with RE = Y, La, Ce, Pr, Nd, Sm and MM. It is obvious
that the SmCos compound has the highest anisotropy field. Figure 1.7 shows
the dependence of anisotropy field, H., on individual rare earth in binary
RE 2C0 17 compounds with RE = Y, Ce, Pr, Nd, and Sm. It is interesting that in
all of the 2 : 17 compounds, only Sm2 C0 17 has uniaxial magneto-crystalline
anisotropy and, therefore, possesses a large crystalline anisotropy field. Sm
even behaves uniquely in ternary RE2Fe14 B compounds. The magneto-
crystalline constant K 1 of Sm2Fe14B is a negative value as shown in Fig. 1.8.
This means that the easy magnetization direction for Sm2 Fe14 B is in the basal
plane rather than along the c-axis, such as in the cases for RE = Y, Ce, Pr,
and Nd. However, the absolute value of K 1for Sm2 Fe14 B is the largest, which
means that the largest magneto-crystalline anisotropy exists between the basal
plain and the c-axis for Sm2 Fe14 B.

1200
,....., 1000
~
;:r, 800
,f
~ 600
& 400
E
~
" 200
'C
::l
U 0

-200 '---:'-:c--=7o=-----=-75=---8-:!:o---:8~5---:9':-0---:9-=-5
~100
Co content (at.%)

Figure 1.5 Curie temperature, Tc , versus Co content in Sm-Co binary compounds.


Data for LaCo13 and Sm2 Co 14 B are also included.
6 Sam (Shiqiang) Liu

500

O'---'---....L------'----'---'''-----'---'---
Y La Ce Pr Nd Sm MM
Rare earth element

Figure 1. 6 Anisotropy field, H., in binary RECos compound.

80
70
o
g" 60
~ SO
-0
~ 40

~ 30
o
E 20
~

Figure 1. 7 Anisotropy field, H., in binary RE2Co 17 compound.

8
o
o

Rare earth element

Figure 1.8 Anisotropy constant K 1 in RE2Fe14B ternary compounds.

It can be concluded from the above analysis that until a totally new high-
temperature permanent magnet material is discovered, an appropriate approach to
develop better high-temperature permanent magnet materials that satisfy the
Recent Developments in High-Temperature Permanent Magnet Materials 7

requirements of the advanced power devices is to significantly improve the


conventional Sm-Co type magnets, especially Sm2 (Co, Fe, Cu, Zr) 17 magnets
and to develop new compounds based on Sm-Co compounds. In order to satisfy
high-temperature (>400 C) applications, the new materials have to possess
high intrinsic coercivity at high temperafures (>400 C) so that the induction
demagnetization curve remains linear at the operating temperature. Under this
condition, the irreversible loss of magnetization caused by high temperature
can be minimized.

1. 3 New Sm2 Co17 -Based High-Temperature Permanent


Magnet Materials

In recent years, extensive research has been carried out to substantially


improve the high-temperature performance of the Sm-TM permanent magnets,
and a breakthrough was made in 1999: As a result of this breakthrough, the
maximum operating temperature of permanent magnets has been increased
from around 300C to as high as 550 C . This advance was made on systematic
studies of the effects of compositions on the high-temperature intrinsic
coercivity of Sm-TM type of permanent magnets.

1. 3. 1 Effects of Compositions on High-Temperature Intrinsic


Coercivity of Smz-TM17 Permanent Magnets

1. 3.1. 1 Effect of Sm on High-Temperature Intrinsic Coercivity of Sm-TM


Permanent Magnets
The Sm content, or the z value, in Sm (Co, Fe, Cu, Zr) z' strongly affects the
intrinsic coercivity of Sm2 (Co, Fe, Cu, Zr) 17 -type of permanent magnets. As
early as 1982, S. Liu et al. (1982) determined the temperature dependence of
a few SmCos and Sm2 TM 17 permanent magnets up to 750 C . They observed
that a magnet specimen of SmTM7 that had a smaller z value of 7 (higher Sm
content) possessed higher intrinsic coercivity than a magnet specimen of
SmTM7 . 43 that had a larger z value of 7.43 at high temperatures. More
recently, Kim ( 1998) observed that a SmTM7 magnet had a much lower
temperature coefficient of coercivity than a SmTMa magnet. With a high Sm
content (low z value) in the magnet, an intrinsic coercivity of 5 kOe was
achieved at 500C. J. F. Liu et al. (1999a, 1999b) studied the high-
temperature magnetic properties of SmTM z with z = 6.7 - 9. 1 at
temperatures up to 500 C . They found that the lower z value led to a smaller
8 Sam (Shiqiang) Liu

temperature coefficient of intrinsic coercivity. When z was changed from 7. 0


to 8. 5, the temperature coefficient of intrinsic coercivity changed from
-0.03%/"C to -0.25%/"C.
When dealing with the effect of Sm content or z value, it is important to
realize that not all Sm exists in the Sm-TM compounds in sintered Sm2 (Co, Fe,
Cu,Zr) 17-type of magnets. A small amount of Sm exists in the form of Sm203'
because Sm is a very active element, and some Sm is oxidized during the fine-
powder processing and oxide formation. Under normal conditions, a sintered
Sm2 (Co, Fe, Cu , Zr) 17 magnet contains O. 3 - O. 6 wt. % oxygen. It is easy to
understand that oxygen reduces the effective Sm content by 6. 27 times the
mass fraction of the oxygen. This means that for each O. 1 wt. % oxygen, there
will be 0.627 wt. % Sm to be consumed and reacted with oxygen. Therefore,
it is useful to define an effective z value that represents the atomic ratio of the
transition metals over the metallic part of Sm. Table 1.2 shows the effective z
in Sm (Coo.sFeo.1Cuo.oSZrO.02ho with differing oxygen contents. These data
were calculated under an assumption that all oxygen in sintered Sm-TM
magnets exists as Sm203' and it is believed that Zr is more likely to react with
C in the alloy and form ZrC rather than ZrO (Liu et aI., 1999a, 1999b) .It can
be seen from Table 1.2 that when the oxygen content increases from 0 wt. %
to 0.60 wt. %, the effective z value changes from 7. 00 to 8. 16. Therefore,
stating a nominal z in rare earth permanent magnets without simultaneously
mentioning the oxygen content in the magnet is not very meaningful. The z
values given in this chapter are the effective z unless otherwise specified.

1. 00
Effective z 9.17

Generally speaking, decreasing the z value (increasing the Sm content) of the


conventional 2 : 17 magnets results in decreased intrinsic coercivity at room
temperature and increased intrinsic coercivity at high temperatures. Figure 1. 9
summarizes the effect of z value on MH c of Sm (COO.795 Feo.09 CUO.09 ZrO.025)z
(Liu et al. ,2000a). Both effective and nominal z values are given in this
figure. The effective z value in most conventional 2: 17 magnets is
approximately 8. 3-a value much higher than all magnets presented in this
figure. It can be seen from Fig. 1. 9 that MH c at room temperature is very
sensitive to the z value and MH c increases rapidly with increasing z. As the
temperature rises, especially when T~500 "C , MH c becomes less sensitive to
z. It can be seen that there is a peak (denoted by pk) in each MHc-z curve in
the temperature range of 300 - 500 "C . It is interesting to note that as the
temperature increases, the z value corresponding to the peak MH c shifts
toward lower values of z. As shown in Fig. 1. 10, the effective z values
corresponding to the peak MHc at 300, 450 and 500 tare 7.86, 7.62 and 7.38,
Recent Developments in High-Temperature Permanent Magnet Materials 9

Nominal z
6.4 6.6 6.8 7.0 7.2 7.4
28 pk ___
1:20C
24 2: 300C
3: 350 C
4: 400 C
g 20 5:450C
6: 500C
C
~ 7: 550C
:E 8:600C
616
';> 2
.~
0
0
0
12 3
.;;;
c:
'5 4
..s 8

oL -_ _..L-_ _---L_ _----1 "---_---.J


7.02 7.26 7.50 7.74 7.98 8.22
Effective z
Figure 1.9 Dependence of intrinsic coercivity of Sm(COo,795 Feo.OgCuO,OgZrO.02S)z
on z value at various temperatures.

18 z=7.86
7.98
Sm (Coo.795Feo.09CUOQ9ZrO,02S)z
16 8.10
V' 7.62
0-'<: 14
;:r:E 7.50
612
';>
.~ 7.38
.,
8 10
0
.;;; 7.26

~ 8 z=7.14

250 300 350 400 450 500 550 600 650


Temperature (OC )

Figure 1. 10 Temperature dependence of intrinsic coercivity for


Sm(COo,795FeO,OgCUO,OgZro,025)zfrom 300 to 600C.
10 Sam (Shiqiang) Liu

respectively. It is obvious that at room temperature, the peak should occur at


z ~8. 10, while at 550 and 600 C , the peaks should occur at z~7. 14.
It can also be concluded from Fig. 1.9 that MH c becomes more and more
sensitive to temperature with increasing z. This trend can be seen more clearly
from Fig. 1. 10, which shows the temperature dependence of MH c for 8m
(Coo.795Feo.ogCuo.OgZro.o2s)z' It can be seen from Fig. 1. 10 that when the
effective z = 7. 14, the coercivity slightly increases with increasing
temperature in the temperature range from 300 to 450C. As z increases, MH c
gradually increases in the temperature range of 300 - 450C and displays
increased negative temperature coefficients of MH c . When z exceeds 7. 86,
MH c decreases with z at the entire temperature range from 300 to 600 C .

1. 3.1. 2 Effect of Fe on High-Temperature MHc of Sm-TM Magnets


Fe also has a strong effect on high-temperature coercivity of 8m2 (Co, Fe, Cu,
Zr) 17 -type permanent magnets. Increasing Fe content in 2 : 17 type magnets
effectively enhances their saturation magnetization and maximum energy
product. However, it was observed in the mid-1990s that high Fe content (or
low Co content) resulted in poor high-temperature stability of the magnets (Liu
and Hoffman, 1996; Ma et aI., 1996; Liu et al. , 1997). In 1998, C. H. Chen
et al. (1998) obtained sintered 8m2 (Co, Fe, Cu, Zr) 17 -type permanent
magnets with improved high-temperature performance by reducing the Fe
content in the magnet alloys. A high intrinsic coercivity of 8. 3 kOe at 400 "C
was achieved when the Fe content was decreased to 7 wt. %. Their 2 : 17
magnets with low Fe content demonstrated lower irreversible loss of magnetic
flux, higher maximum energy product, and lower temperature coefficient of
intrinsic coercivity as compared with the conventional 2 : 17 magnets that
contain 15 wt. % - 20 wt. % Fe. In 1999, J. F. Liu (1999a, b) accomplished
intrinsic coercivity of 10.7 kOe at 500 "C in a cast magnet alloy.
Generally speaking, decreasing Fe content (increasing Co content) in the
2 : 17 magnets does not strongly affect their room temperature coercivity, but
significantly enhances their coercivity at high temperatures. The effect of Fe
content on intrinsic coercivity in 8m (Coba,Fe.Cuo.OgZro.o3 hs at high
temperatures from 400 to 600C is given in Fig. 1. 11 (Liu 8. et al. , 2000c).
It should be noted that the Fe conterit in conventional 2 : 17 magnets is v =
O. 21 - O. 31, much higher than the Fe content in magnet materials in this figure. It
can be seen from Fig. 1. 11 that the high-temperature MH c increases rapidly
with decreasing Fe content in the magnet alloys. At 400 "C , MH c increases from
5.6 kOe when v = 0.22 to 12.7 kOe when v = O. 1. A coercivity peak
appears when v = O. 1 at 400 "C . This peak shifts to v = O. 07 at higher
temperatures.
Recent Developments in High-Temperature Permanent Magnet Materials 11

14

12

~
<l)
0 10
6
~
:2 8
is
.:;
.~
<l)
0
()
6
()
'v;
c:
.;:
. 4

0 0.05 0.10 0.15 0.20 0.25


Fe content, v

Figure 1. 11 High-temperature intrinsic coercivity, MH e , as a function of Fe content, v,


in Sm(Coba,Fe.Cuo,OgZrO,o3)7.5.

1. 3. 1. 3 Effect of Cu on High-Temperature MHc of Sm-TM Magnets


It is well known that coercivity in the Sm2 (Co, Fe, Cu, Zr) 17 type of magnets
originates from the pinning of the domain wall in the Cu-rich cell boundary
phase in a fine-scaled cellular microstructure (Livingston and Martin, 1977;
Mishra et aI., 1981; Ray et aI., 1987). Therefore, sufficient Cu content is
essential to develop high coercivity at both room temperature and high
temperatures. Generally speaking, MH e increases with the Cu content
monotonously. However, the effect of Cu on increasing MH e is quite different at
different temperatures. Figure 1. 12 shows the temperature dependence of
intrinsic coercivity for Sm(Co ba, Feo, 1 CuxZrO.025 )6,7' It was observed that when a
magnet contains very low Cu, its MH e could have a positive temperature
coefficient. This abnormal temperature dependence of intrinsic coercivity is
explained in Section 1.3.6.
1. 3.1. 4 Effect of Zr on High-Temperature MHc of Sm-TM Magnets
Zr has an important effect on the development of coercivity in the Sm2 (Co, Fe,
Cu, Zr) 17 -type magnets. It has been observed that Zr is critical in developing
high coercivity at both low and high temperatures. It was observed that
intrinsic coercivity rapidly increased with increasing Zr and a peak coercivity
value was reached at an optimum Zr content. The squareness of the 2nd
quadrant demagnetization curve is strongly depended on Zr content in magnet
alloys. The knee field is rapidly enhanced with increasing Zr content.
12 Sam (Shiqiang) Liu

16

x=0.05

o 100 200 300 400 500 600


Temperature ('C )

Figure 1. 12 Temperature dependence of intrinsic coercivity for Sm( C~l Feo. 1 Cu. ZrO.025)6.7 .

Figure 1. 13 shows the effect of the Zr content on the high-temperature intrinsic


coercivity in 8mCCoba,Feo.osCuo.oaZry)7.62 .It is obvious that when the Zr content
is lower than y = O. 01, the coercivity is very low at all temperatures. The
coercivity is substantially enhanced when the Zr content increases to
y = O. 018. It is also seen from this figure that coercivity peaks appear for
300, 350, 400 and 450C when y = O. 026. As temperature increases higher
than 450 C , magnets containing more Zr display sl ightly higher coercivity.
16
14
~

'"
~12
~10
~
.;;: 8
.~

8'" 6
o
.~ 4

~ 2

o 0.005 0.010 0.0 15 0.020 0.025 0.030 0.035


Zr content, y

Figure 1. 13 Effect of Zr content on high-temperature intrinsic coercivity in


Sm(C~IFeo.09Cu008ZrY)7.62 .

The effects of other transition metals such as Ti, Hf, Nb, V, Ta, Cr and
Ni on the high-temperature coercivity of the 8m2 CCo, Fe, Cu, Zr) 17 were also
investigated. All these elements decreased magnetization. Only Nb
demonstrated the effect of slightly enhanced coercivity at high temperatures.
Recent Developments in HighTemperature Permanent Magnet Materials 13

1. 3. 2 Magnetic Properties of New 8m2 Co17 -Based High-Thmperature


Permanent Magnet Materials
The above-mentioned research on the effects of compositions on high-
temperature performance of the 2 : 17 type of permanent magnets has resulted
in a new series of sintered permanent magnets with significantly improved
high-temperature performance accomplished by significantly reducing the Fe
content, increasing the 8m content, and adjusting the Cu and Zr contents in the
magnet alloys.
The operating temperature of these magnets has been increased from the
previous 300C for conventional high-temperature magnets to as high as
550 C . The MH c of these new magnets reached 13 kOe at 400C (two to three
times higher than conventional magnets) and 9 kOe at 500C (four to nine
times higher than conventional magnets). The B curves of the new magnets
remain linear up to 550 "C (250 to 350C higher than conventional magnets).
The temperature coefficients of MH c for the new magnets can range from a
small negative value ( - O. 03 % FC ), to near zero, or they may even be
positive (up to + 0.3 % FC ) . As a comparison, the temperature coefficients of
MH c for conventional 8mCos , 8m2 TM 17 and Nd2 Fe14 B magnets around room
temperature are - O. 3 % FC, - 0 . 3 % /"C and - 0 . 9 % FC , respectively.
Figure 1.14 shows the demagnetization curves of 8m(Cot,a,Feo.o9Cuo.o9Zro.o3)7.69
at 400C , 450C and 500C (Liu et al. , 1999a). This magnet illustrates much
higher MH c and better squareness of demagnetization curves at high
temperatures than the conventional 2 : 17 magnets. Figure 1. 15 compares
induction demagnetization curves at high temperatures up to 550C for the best
conventional high-temperature 2 : 17 magnet and the new high-temperature
magnets (Walmer et al. , 2000b). It can be seen from Fig. 1. 15 that the best
\2

glO

16 6
8
41tM

<::
o
.~ 4
N
.~

~ 2
0
::E
_\L-_--:'(L--'-_-'-\-=-2__ll.(O:---_~8....L.<:...<-L6---'-4:-----'=2:---
6 4
Magnetic field strength (kOe)

Figure 1.14 Demagnetization curves of Sm(Co".,FeO.09CuO.09ZrO.03 )7.69


at 400 t , 450C and 500 t .
14 Sam (Shiqiang) Liu

12

MagnetB: TM =330"C
10
0-
~
'Cl 8
Load line
'"
0
.~
;:l
"0
6
L/d=0.5

.::
.g
4
'"6b
:2'"
2

0
-12 -8 -6 -4 -2
Applied magnetic field, H (kOe)
(a)
12

Demagnetization curves at each TM A


10
0- B
~ C
'Cl 8
A: 250"C, 25.4 MG Oe
'u'"
0
B: 330"C, 20.1 MGOe D
;:l 6 C: 400"C, 16.5 MGOe
"0 E
.:: D: 500"C, 10.4 MGOe
'"
.~
4
E: 550"C , 6.3 MGOe
c
on
:2'"
2

0'--_--l.,...--L-...Ll..-...L-_...1.L_...L-.L-_-----.J'-----_----'
-12 -10 -8 -6 -4 -2
Applied magnetic field, H (kOe)
(b)

Figure 1. 15 Induction demagnetization curves at high temperatures up to 550"0 for


conventional 2 : 17 (a) and new high-temperature magnets (b).

conventional high-temperature 2: 17 magnet remains a linear induction


demagnetization curve up to 330 C . At 400 C , the induction demagnetization
curve is severely bent and a knee appears, while the induction
demagnetization curves for the new high-temperature magnets remain linear up
to 550 C. Figure 1. 16 shows demagnetization curves for a new high-
temperature magnet with its maximum operating temperature T M 500 C . =
These new magnets are now commercially available from electron energy
corporation.
The temperature dependence of MH c for some newly-developed sintered
8m-Co based permanent magnets is shown in Fig. 1. 17 (Liu et al. , 2000b).
For comparison, the temperature dependence of MH c of a high-coercivity type
of conventional 2 : 17 magnet is also shown as Curve 1 in the figure. It can be
Recent Developments in High-Temperature Permanent Magnet Materials 15

Magnet D (TM=500 C ) demagnetization curves,.f(T)

Temp. B, Hei He (BH)max Temp. B, Hei He (BH)max 14


CC) (kG) (kOe) (kOe)(MGOe) CC) (kG) (kOe) (kOe)(MGOe)
25 9.28 28.60 8.92 20.84 300 8.06 16.72 7.50 15.24
100 9.06 25.67 8.63 19.68 350 7.81 14.55 7.23 14.31 12
150 8.88 23.72 8.42 18.80 400 7.55 12.37 6.91 13.17
200 8.64 21.56 8.15 17.73 450 7.20 9.81 6.40 11.84
250 8.40 19.17 7.89 16.68 500 6.88 7.32 5.63 10.43 10 8
~
...
0

"
~
~
-0
c:
~
....l':

-26 -24 -22 -20 -18 -16 -14 -12 -10 -8 -6 --4 -2

-2000 -1600 -1200 -800 --400 o kNm


Magnetic field, H (kOe or kAhn)

Figure L 16 Demagnetization curves of a new high-temperature magnet with T.. = 500 0 .

<>: Conventional 2: 17
30 0: Sm (CObalFeo 09CUo.Q9Zro.oJh69
v: Sm (CObaIFeo09CUo.OgZrO.02S)769
,-., 0: Sm (CobaIFeo04CuO.IZr003h38
Q)
0 25 t.: Sm (CObaIFeo.09CU0.Q9ZrO.02Sh.14
~
x: Sm (CobaIFeo04CuO.09ZrO.027)726
:zr::;:
20
~
.;;
.~
Q)
0
C,)
15
C,)
'Vi
c:
'5 10
..5

0 100 200 300 400 500 600


Temperature CC )

Figure L 17 Temperature dependence of intrinsic coercivity of various


Sm(Oo,Fe,Ou,Zr)z magnets.
16 Sam (Shiqiang) Liu

seen from Fig. 1. 15 that the MH c of the new magnets is much less temperature
sensitive in comparison with the conventional 2 : 17 magnet. At temperatures
above 100C, the MH c of Sm(CobaIFeo.o9Cuoo9Zro.o3)7.69 (Curve 2) is higher
than that of the conventional 2 : 17 magnet. At 400 C , the MH c is three times
higher than that of the conventional magnet. Curve 4 has a very flat portion at a
quite high coercivity level from 200 to 400 C . From room temperature to
450 C, the MH c of Sm (CObal Feo.o9 CUO.09 ZrO.025 h 14 remains almost constant
(Curve 5), making its temperature coefficient very close to zero over this
wide temperature range. In addition, a complex temperature dependence of
MH c in some newly-developed magnets was observed (Curve 6). Further
details will be presented in Section 1.3.6.2.
The variation of temperature coefficients of MH c as a function of
temperature for the magnets shown in Fig. 1. 15 is given in Fig. 1. 18 (Liu et
al. , 2000b). As a comparison, Fig. 1. 18 also gives temperature coefficients of
MH c of a typical SmC05 and a Nd-Fe-B magnet. The temperature coefficient of
MH c demonstrated in Fig. 1. 18 is defined as (Liu et aI., 1990b)

/3=d(MH c )x 100 (%;oC). (1. 1)


dT MH c
Detailed information about this definition and its calculation will be given in
Section 1. 4.
0.4

~ 0.2
~

e~Q::; -0.2O~i~~~~;;::~F=:~
I'
'-
o
C -0.4

'"lE" -0.6
8"
" -0.8
5
~ -1.0
0-
~ -1.2
f-
-1 .4 l-----L-------''-''------'-----'-_--'---'-_-'----'-_-'---'-_''--'-'L---l
o 50 100 150 200 250 300 350 400 450 500 550 600 650
Temperature ('C )

Figure 1. 18 Temperature coefficients of various permanent magnets.


Numbers (1- 6) denote the same magnets as shown in Fig. 1. 17.

1. 3. 3 Dynamic Characterizations of New High-Temperature


Permanent Magnets
It is important to realize that the typical magnetic characterization is static. Its
results are only meaningful to the static applications in which the operating
Recent Developments in High-Temperature Permanent Magnet Materials 17

point of a magnet in the induction demagnetization curve is fixed. Data obtained


from a static characterization do not necessarily represent the real capability
of a magnet in dynamic applications in which the operating point of the magnet
in the induction demagnetization curve is cycled. Examples of dynamic
applications are motors. generators. magnetic bearings. and magnetic actuators.
Because these important applications of permanent magnets are dynamic.
it is of the utmost importance to know the capability of a magnet under dynamic.
rather than static conditions. Bearing this in mind. we devised two sets of dynamic
characterizations. In the first one. the magnet specimens were magnetically cycled
under an applied demagnetizing field of 0 to sHe (s He denotes induction coercivity)
before their magnetic properties were determined. The first set of characterizations
was performed from room temperature to 500 C. The second set of
characterizations was carried out at a fixed temperature of 400 C with an
applied demagnetizing field ranging from 0 to 9 kOe. We compared the
dynamic magnetic properties of the new magnets with those of the best
conventional 2 : 17 magnets in these two sets of dynamic characterizations. The
maximum operating temperature of the magnet specimens used in these dynamic
characterizations is 450 C .
Results of the first set of dynamic characterization are given in Figs. 1. 19
through 1.22 and summarized in Fig. 1.23 (Liu et al . 2000c). It can be seen
from these figures that at room temperature. the conventional 2 : 17 magnets
have higher dynamic (BH)max' However. at about 250C. the dynamic
(BH)max of the conventional 2 : 17 magnets begins to sharply decrease. At 400
C its (BH) max drops to less than 1/20 of the new magnet as shown in
Fig. 1.21. Results of the second set of dynamic characterizations are
summarized in Fig. 1.24 (Liu et al. 2000c). It is obvious that the maximum
applied demagnetizing field to which the best conventional 2 : 17 magnet can
be subjected without significant loss of (BH)max is about 2 kOe. while for the
new magnet. it is as high as 8 kOe.
16
Dynamic characterization at 20C
14 (magnet cycled between 0 and SHe)
6' I: Best conventional 2: 17 magnet
012
c:: 2: New 8m-Co based magnet
o
.~ 10
] 2
<; 8
c::
o 6
:~
I 4
2

0'-----1.--'---'--'--------"'---"--'---'--'--------'--'
-20 -18 -16 -14 -12 -10 -8 --{i -4 -2 0
Magnetic field strength (kOe)

Figure 1. 19 Result of dynamic characterization performed at 20 0 .


18 Sam (Shiqiang) Liu

16
Dynamic characterization at 300'C
~ 14 (magnet cycled between 0 and sHe)
g
12 1: Best conventional 2: 17 magnet
.~ 2: New Sm-Co based magnet
}1 10
6 8
c
0

:~ 6
"6iJ
oj 4
~
2

o"--_"---u_---'_-'-_--'-_-J...<.---'._-"-_-'-----'
-20 -18 -16 -14 -12 -10 -8 -{; -4 -2 0
Magnetic field strength (kOe)

Figure 1. 20 Result of dynamic characterization performed at 300 '0 .

16
Dynamic characterization at 400 'C
~14
(magnet cycled between 0 and sHe)
g
c 12
I: Best conventional 2: 17 magnet
"B:> 2: New Sm-Co based magnet
"t:l 10
.5
6 8
c 2
.9
.~c 6

~ 4
~
2

o"--_"-----'_---'_.L...L_--'-_-J...---""-'-_-"-_-'-L...-J
-20 -18 -16 -14 -12 -10 -8 -{; -4 -2 0
Magnetic field strength (kOe)

Figure 1. 21 Result of dynamic characterization performed at 400 '0 .

16
Dynamic characterization at 450'C
G' 14 (magnet cycled between 0 and sHe)
C
c 12
0
fl 1: Best conventional 2: 17 magnet
:> 2: New Sm-Co based magnet
10
..
"t:l
.:
0
c 8
0
.~ 2
6
.~

@,
oj 4
~
2

O'--_"-----''----'_---''_-LL_--'-_-I<.-'--'-_-'--L..J
-20 -18 -16 -14 -12 -10 -8 -{; -4 -2 0
Magnetic field strength (kOe)

Figure 1. 22 Result of dynamic characterization performed at 450 '0 .


Recent Developments in High-Temperature Permanent Magnet Materials 19

30
0: Dynamic characterization for
25 the best conventional 2: 17
'1: Dynamic characterization for
new Sm-Co based magnet
OJ 20
0
0
:z:
~ 15
~
E

~ 10
5

0 100 200 300 400 500


Temperature CC )
Figure 1. 23 A summary of the first set of dynamic characterizations.

20
18
Dynamic Characterization at 400C
16
14 2
.,
8 12
~ 10

~ 8
6
4
0: Best conventional 2: 17
2 0: New Sm-Co based magnet

o 23456789 10
Maximum negative cycling field (kOe)

Figure 1. 24 A summary of the second set of dynamic characterizations.

1. 3. 4 Long-Term Thermal Stability of New High-Temperature


Permanent Magnets
An aging experiment was performed to test the long-term high-temperature
stability of the new magnets. Figure. 1.25 shows the losses of flux density in
the first 100 h for 8m (COO.795 Feo.o9 CUO.09 ZrO.025) z' with effective z = 7.26,
7. 62, 7. 86 and 8. 10, in a conventional long-term aging experiment performed
at 500C in air. As comparisons, data for a best conventional high-temperature
2 : 17 - 28 magnet (room temperature (BH)mex = 28 MGOe) , a conventional
20 Sam CShiqiang) Liu

2 : 17-30 magnet (room temperature (BH)max= 30 MGOe) are also included


in the figure. As can be seen from the figure, after aging at 500C for 100 h,
the flux density losses are about 56 % and 36 % for the conventional high-
temperature 2 : 17 - 30 and 2 : 17 - 28, respectively. While under the same
condition the losses for the best new magnet is only 3% (Liu et al. , 2000c).
Figure 1. 26 shows the loss of flux density up to 2000 h for the same
magnets. It can be seen from Fig. 1.26 that after aging at 500 C for 2000 h,
the flux density losses are 65 % and 49 % for the conventional high-temperature
2 : 17 - 30 and 2 : 17 - 28, respectively. Under the same condition, the losses
for the best new magnet is 19% (Liu et al. , 2000a,2000c).

z 7.62
z 7.86
z 8.1
z-7.26
ConventionaI2:17-28
(best pre-project magnet)

Conventional 2: 17-30

-70
500 C air aging
-80'------'-----'------'-----'------
o 20 40 60 80 100
Aging time at 500C (h)

Figure 1. 25 Loss of flux density vs. aging time at 5000 for the first 100 h.

0 500C air aging


-10
~
z=7.62
-20
~ z=8.10
0
c: -30
'iii z=7.86
'"
""Cl z=7.26
><:
....... --40
:::l

a
-50
'"'"a
....l Conventional 2: 17 - 30
-<>0

-70

-80
0 500 1000 1500 2000
Aging time at 500C (h)

Figure 1.26 Loss of flux density vs. aging time at 5000 up to 2000 h.
Recent Developments in High-Temperature Permanent Magnet Materials 21

It is obvious from Fig. 1.26 that the newly developed magnets display
significantly lower losses of flux density than the conventional 2 : 17 magnets
under the same testing condition. The magnet with z = 7. 62 gives the best
long-term stability. Magnets with the z values lower or higher than 7. 62
illustrate relatively larger losses. It is observed that the magnet with z = 7.26
displays the largest loss among the four new magnets though at high
temperatures it has higher MH c than magnets with z = 7. 86 and z = 8. 10
(Fig. 1.9). This is related to the fact that it has low room temperature MH c .
A recent study has revealed that the flux density loss of the best high-
temperature magnet during the high-temperature in-air aging as shown in
Figs. 1.25 and 1.26 is caused by the diffusion of Sm to the specimen surface
rather than by metallurgical change (Walmer et al. , 2000b). Therefore, the
loss can be further reduced by applying a proper coating layer.

1. 3. 5 Microstructure and Crystal Structure

Figure 1.27 is an optical micrograph of Sm(CoO.794 Fe O.lCUO.09ZrO.026ho after


aging. The nominal z value of the magnet specimen is 6. 46. In other words,
the magnet specimen contains very high Sm. Under this condition, it still shows
a uniform" one-phase" microstructure. The fine-scaled cellular microstructure
of high-temperature magnets has been observed by many researchers. Kim
(1998) first observed that the cell size is smaller in a SmTM7 magnet than in a
SmTMs magnet. Similar results were reported by. Liu et al. (1999a), Schrefl
et al. (2000), and Walmer et al. (2000a, 2000b). Figure. 1. 28 is a transmission
electron microscopy (TEM) micrograph of a new high-temperature magnet.

Figure 1. 27 Optical micrograph of Sm(COo.794FeO.l ClJo.OgZr 0.026 )7.0 (Nominal z = 6.46).


22 Sam (Shiqiang) Liu

Figure 1. 28 TEM micrograph of a new high-temperature magnet.

The X-ray diffraction (XRD) pattern of Sm(CoO. 794 Feo.o9Cuo.o9Zro.o26)7.o


after solid-solution heat treatment (SSHT) is given in Fig. 1.29. This figure
suggest that the magnet specimen may have a 1 : 7 hexagonal crystal structure
in the SSHT condition. Figure 1. 30 is the XRD pattern of the same magnet after
aging and in a high coercivity condition. The XRD result indicates that the
magnet specimen consists of a 2 : 17 rhombohedral phase and a 1 : 5 hexagonal
phase. More accurate determination of the crystal structures would rely on
additional analyses, including neutron diffraction and Mossbauer work.

:::::-
8000
-
:::. 1:7 (Hex.)

~ 6000
tl 80
c::
::l
0 c:!-
~ :::::- N ~

Co 4000 0
8 0 0
.;;; :::. S c:!-
c::
~
.::
2000

40.0 50.0 60.0 70.0 80.0


2en
Figure 1. 29 XRD pattern of Sm(COo. 794 FeO.09 CUO.09 ZrO.026 ho
after solid solution heat treatment.
Recent Developments in High-Temperature Permanent Magnet Materials 23

7000
2:17R+I:5 Hex
6000

~ 5000
c
::>
o ,-....--.. --..-..
84000 --..O"'T_ '0,-.",("<")-..
0 0 0 0 O~N
o
iii
-MC"\lO OONO
.- '-''-' N ' - ' 0 '-'(',1
'-' r---- r- '-' \0 '-' r- '-'
l:5 3000 l r ' l - ......
..:.:~0i.:.:
V') ...... V) ...... V)
r-i":":N':';
:5
2000

30 40 50 60 70 80
2en
Figure 1. 30 XRD pattern of Sm(COo 794 Feo.OgCUO,OgZrO,026 )70 after aging.

1. 3. 6 Coercivity in 8m2 TM 17 -Type Permanent Magnets

1. 3. 6. 1 Explanation of Effects of 8m, Fe, Co on Intrinsic Coercivity


During the development of the new high-temperature permanent magnets, it
was observed that Fe, 8m and Cu contents significantly affect high-
temperature coercivity of 8m2 (Co, Fe, Cu, Zr) 17-type permanent magnets.
Generally speaking, decreasing Fe (increasing Co) content does not strongly
affect room temperature coercivity but leads to a significantly higher coercivity
at high temperatures; increasing 8m content results in much lower room
temperature coercivity but substantially higher coercivity at high
temperatures; and increasing Cu content leads to higher coercivity at all
temperatures. The effects of Fe, 8m and Cu on coercivity can be explained by
using the temperature dependence of magneto-crystalline anisotropy of the 2 :
17 cell phase and 1 : 5 cell boundary phase.
( 1) Effects of changing Fe, 8m and Cu Content on Element
Concentrations in the Cellular Microstructure
It is well known that coercivity in 2 : 17 magnets originates from the
domain wall energy difference between a 1 : 5 cell boundary phase and a 2 :
17 cell phase in a fine-scaled cellular microstructure (Livingston and Martin,
1977; Mishra et aI., 1981; Ray et aI., 1987). The 1 : 5 cell boundary phase
has lower crystalline anisotropy K and, in turn, lower domain wall energy E
than the 2 : 17 cell phase. Therefore, domain walls are preferably pinned in
the 1 : 5 cell boundary phase.
Figure 1.31 gives composition maps for a typical high-temperature 2 : 17
magnet (Liu et al. , 2001). It shows a bright field image and concentrations of
24 Sam (Shiqiang) Liu

Zr, Co, Fe, Sm, and Cu, respectively. In the composition maps, a bright
region represents a high concentration of an element. It can be clearly seen
from Figure 1. 31 that the cell interior is rich in Co and Fe, while the cell
boundary phase is rich in Sm and Cu.

Figure 1. 31 Composition map of a high-temperature Sm2 (Co, Fe, Cu, Zr) 17 magnet.

The Fe content in the 1 : 5 cell boundary phase of 2 : 17 magnets


containing high Fe content is less than 5 at. % (Ray, 1990). In contrast, Fe
substitution for Co can be made up to 100 at. % of Fe in the 2 : 17 phase (Ray
and Strnat, 1972). Therefore, it is assumed that changing the Fe content in the
2 : 17 magnets primarily alters the composition of the 2 : 17 cell phase. On the
other hand, considering the facts that: CD the SmCos phase has a relatively
Recent Developments in High-Temperature Permanent Magnet Materials 25

high solubility for 8m even at low temperature 400 'C) and can be
expressed as 8mC05 - 6 with {) > 0 (Khan, 1974), which is different from
8m 2 Co 17 ; <?> most of the Cu in 2 : 17 magnets is concentrated in the 1 : 5 cell
boundary phase (Ray, 1990); and @ Cu substitution for Co in 8mC05 can be
made up to 100 at. % of Cu (Ray, 1990). It is assumed that changing the
contents of 8m and Cu primarily affects the composition of the 1 : 5 cell
boundary phase.
(2) Effect of Decreasing Fe (Increasing Co) Content
Perkins et al. (1975) showed that when the Fe substitution for Co was
decreased from 40 at. % to 20 at. % , the room temperature anisotropy field of
a 2 : 17 alloy rapidly increased from 32 to 65 kOe. However, when the Fe
content was further decreased from 20 at. % to Oat. %, the anisotropy field
did not show a significant change. In conventional 2 : 17 magnets, the Fe
content is around 20 at % . The Fe content has been decreased to 0 at. % - 10
at. % in the newly developed high-temperature 2 : 17 magnets. Therefore, it
can be concluded that decreasing Fe content does not lead to a significant
change of the crystalline anisotropy in the 2: 17 cell phase at room
temperature (T nn)' This is shown in Fig. 1.32 which schematically illustrates
the temperature dependence of crystalline anisotropy for the 2 : 17 cell phase
(Curve 1) and the 1 : 5 cell boundary phase (Curve 2) in a conventional 2 : 17
magnet and for the 2 : 17 cell phase in a 2 : 17 magnet containing low Fe (Curve 3).

I: 2:17 cell phase


2: 1:5 cell boundary phase
3: 2: 17 cell phase of magnet containing low Fe

..... .....
......
...... ......
...... .......
........

Temperature

Figure 1. 32 Schematic temperature dependence of magneto-crystalline anisotropy for the


2 : 17 cell phase and 1 : 5 cell boundary phase in a conventional 2 : 17 magnet and for the
2 : 17 cell phase in a 2 : 17 magnet containing low Fe.

On the other hand, decreasing Fe content (increasing Co content) results


in a significant increase of Curie temperature of the 2 : 17 cell phase (Ray and
8trnat, 1972). Thus, Curve 3 should be higher than Curve 1 in the high
temperature region as shown in Fig. 1. 32. This explains why decreasing Fe
content does not strongly affect room temperature coercivity, but leads to a
significantly higher coercivity at high temperatures.
26 Sam (Shiqiang) Liu

(3) Effect of Increasing 8m Content


As discussed earlier, when the 8m content in the magnet alloy is
increased, the extra 8m primarily enters the cell boundary region. 8ince the
magneto-crystalline anisotropy in 8m-Co compounds mainly originates from the
8m-8m sublattice , it is reasonable to believe that increasing the 8m content in
8m(Co, CU)5-6 would result in an increased crystalline anisotropy at room
temperature, T rm' and at low temperatures.
It is well known that the magnitude of the Curie temperature in 8m-Co
compounds is determined by the Co-Co sublattice. Thus, increasing 8m
content in 8m(Co,Cu)5-6 leads to a decreased Curie temperature of the 1 : 5
cell boundary phase. From the above analysis, the temperature dependence of
the crystalline anisotropy of the 1 : 5 cell boundary phase in the magnet
containing high 8m would be a Curve 3 shown in Fig. 1.33. It is obvious that at
room temperature and low temperature, the magnet containing high 8m
displays lower coercivity, while at high temperature, it displays higher
coercivity as compared with conventional 2 : 17 magnets.

'< I: 2:17 cell phase


2: 1:5 cell boundary phase
~ 3: 1:5 cell boundary phase of magnet containing high Sm
8
~
15o
U)

'
Q)

:5
"3
~
(j

3 1
i
~
2
-3
Temperature

Figure 1. 33 Schematic temperature dependence of magneto-crystalline anisotropy


constants for the 2 : 17 cell phase and the 1 : 5 cell boundary phase in a conventional
2 : 17 magnet and for 1 : 5 cell boundary phase in a 2 : 17 magnet containing high Sm.

(4) Effect of Increasing Cu Content


When the Cu content in 2 : 17 magnets is increased, the extra Cu mainly
goes to the 1 : 5 cell boundary 8m ( Co, Cu) 5- 6 phase, resulting in both a
decreased crystalline anisotropy and a decreased Curie temperature (Lecrard
et al., 1994). This results in a Curve 3 of the temperature dependence of
crystalline anisotropy of the 1 : 5 cell boundary phase shown in Fig. 1. 34. It is
then understandable that the magnet containing higher Cu shows higher
coercivity at all temperature ranges.
Recent Developments in High-Temperature Permanent Magnet Materials 27

1: 2:17 cell phase


2: 1: 5 cell boundary phase
3: 1:5 cell boundary phase of magnet containing high Cu

2
3
Temperature

Figure 1. 34 Schematic temperature dependence of magneto-crystalline anisotropy


constants for the 2 : 17 cell phase and the 1 : 5 cell boundary phase in a conventional
2 : 17 magnet and for the 1 : 5 cell boundary phase in a 2 : 17 magnet containing high Cu.

1. 3. 6. 2 Novel Temperature Dependence of Intrinsic Coercivity


Novel temperature dependence of MH c was observed during the research on
high-temperature permanent magnets in some newly-developed magnets. A
positive temperature coefficient of intrinsic coercivity in SmTM z with z = 7
was reported by Liu et a!. , C1998) . In 1999, Liu et a!. C1999a) observed a
complex temperature coefficient in SmCCoba,Feo.04Cuo.OgZr 0.027)7.26 that had a
low Fe content and a high Cu content. When heating this magnet, MH c first
gradually decreases and reaches a minimum at about 150'C as shown in Fig. 1. 35.
18

20':----:-10:-:0:--~20:-:0:--~30:'c0:------:4:'c00::------::5-=-00::-----.:".600
Temperature (C )

Figure 1. 35 Temperature dependence of magnetic properties of


Sm (CCbaI Feo,04CUo.09 ZrO.027 )7.26.
28 Sam (Shiqiang) Liu

With continued heating, the MH c of this magnet rapidly increases and forms a
maximum at 500 C . The MH c of this magnet at 500C is more than 30 % higher
than its room temperature value. This abnormal temperature variation of MH c
also affects its (BH)max as shown in Fig. 1. 36. Another magnet of 8m
(Coo.825Feo.1Cuo.o5ZrO.025)7.38 that has low Cu content displays a maximum MHc at
550 C , which is nearly four times higher than its room temperature coercivity value
as shown in Fig. 1. 32. As early as 1990, Russian researchers observed the similar
abnormal temperature dependence of coercivity (Popov et aI., 1990a, 1990b).
4

o 100 200 300 400 500 600 700


Temperature ("C)

Figure 10 36 Temperature dependence of intrinsic coercivity of


Sm(COo. 825 Feo.1 CUo.OSZrO.02S )7.38 magnet.

The novel temperature dependencies of MH c observed in the new high-


temperature magnets as shown in Figs. 1.35,1.36 and in Fig. 1. 17 (especially
magnets 5 and 6) challenge previous coercivity theories. In these theories, the
effect of temperature on coercivity is related only to the basic magnetic
parameters, such as saturation magnetization and crystalline anisotropy and,
thus, the novel temperature dependence of MH c can not be explained.
Therefore, a new coercivity model is required and this model will certainly lead to a
better understanding of the coercivity origin in rare earth permanent magnets.
Coercivity mechanisms in permanent magnets are very complicated and
many factors affect the coercivity development. However, the complex
temperature dependence of MH c shown in Fig. 1.35 and 1.36 clearly suggests
that there must be two conflicting coercivity mechanisms that control the
coercivity of the magnet. The first mechanism results in higher coercivity at
higher temperature, while the second mechanism leads to lower coercivity at
high temperature. As a result of these two conflicting mechanisms, a more or
less U shaped MH c versus T curve forms at 20 - 500 C .
(1) Mechanism I -Domain Wall Pinning
As mentioned previously, the coercivity in 2 : 17 type magnets originates
from domain wall pinning in the 8m- and Cu-rich 1 : 5 cell boundary phase of a
Recent Developments in High-Temperature Permanent Magnet Materials 29

fine-scaled cellular microstructure. This is because the cell boundary phase has
a lower crystalline anisotropy constant K and, in turn, lower domain wall
energy E than the 2 : 17 cell phase. The magnitude of coercivity is determined
by the domain wall energy difference between the 1 : 5 and 2 : 17 phases.
Therefore, higher coercivity at higher temperature implies that the energy
difference between the two phases increases with increasing temperature.
Because the Sm- and Cu-rich 1 : 5 cell boundary phase has not only a
lower crystalline anisotropy, but also a lower Curie temperature than the 2 :
17 cell phase, the crystalline anisotropy of the 1 : 5 phase may decrease more
rapidly upon heating than the 2: 17 phase, which may result in the
temperature dependencies of crystalline anisotropy of the 1 : 5 and 2 : 17
phases resembling the curves in Fig. 1.37 .It can be seen from Fig. 1.37 that as
temperature increases, the anisotropy difference between the two phases also
increases. This leads to a higher MH c at higher temperature. It should be noted
that while the pinning mechanism may lead to higher coercivity at higher
temperatures, it may also lead to lower coercivity at higher temperatures
depending on the individual temperature dependence of the 2 : 17 cell phase
and the 1 : 5 cell boundary phase.

Anisotropy difference between


2: 17 and 1:5 phases

Temperature

Figure 1. 37 Schematic temperature dependence of crystalline anisotropy


of 2 : 17 cell phase and 1 : 5 cell boundary phase.

(2) Mechanism II -Thermal Activation of Domain Walls


Since pinning is responsible for higher MH c at higher temperature, the
second coercivity mechanism that leads to lower MH c at higher temperature has
to be related to a process other than pinning. It is postulated that the second
mechanism is closely related to a thermal activation process. This assumption
is based on experimental data obtained in 1981 (Liu et al. , 1982) when it was
observed that the temperature dependence of a 2 : 17 magnet could be exactly
fitted to a simple exponential equation when the temperature was lower than
900 K. It is bel ieved that this was not an accident, but revealed the nature of
the coercivity mechanism in 2 : 17 type magnets.
30 Sam (Shiqiang) Liu

As mentioned previously, coercivity in 2 : 17 type magnets originates


from domain wall pinning in the 1 : 5 cell boundary phase of the cellular
microstructure. In order for the domain walls to break away from the 1 : 5 cell
boundary phase, they must obtain enough energy to overcome the energy
barrier between the 1 : 5 and 2 : 17 phases, namely, IlE . This energy can be
acquired not only from the applied magnetic field Cmagnetic energy), but also
from the thermal energy. It is easy to understand that the higher the
temperature, the higher the thermal energy will be; and the domain walls that
are in a thermal excited condition would have more chance to climb up the
energy barrier. Assuming that the probabilityp of a domain wall to climb up an
energy barrier IlE at temperature T follows the Arrhenius equation,
p= 0 expC-IlE/k a T) C1. 2)
where 0 is a constant and k a is the Boltzmann constant. Apparently, high p
means low coercivity. Thus, the relationship between MH c and temperature T
can be expressed as
C1. 3)
where C is a constant, and the energy barrier is written as IlE CT) to
emphasize that it is a function of temperature. It is apparent that IlE CT) is the
domain wall energy difference between the 2 : 17 and 1 : 5 phases and it can
be written as

C1. 4)
where A 2 : 17 and AI: 5 are exchange constants; K 2 : 17 and K 1 : 5 are crystall ine
anisotropy constants for the 2 : 17 and 1 : 5 phases, respectively; and Oland
O2 are constants. Because the exchange constant A of a ferromagnetic
material is a function of its Curie temperature, T C' it can be written as
C1. 5)
where k a is Boltzmann constant and c is a constant. Substituting Eq. C1.4) into
Eq. C1.3) and redefining the constants gives

C1. 6)
Substituting Eq. C1. 6) into Eq. C1. 3), and considering that K is a function of
T, we have

MH c = c ex p (C1 Jk aCTc )2:17 K CT)2:17- C2 Jk aCTc)l:sKCThs).


kaT
C1. 7)
In a phenomenological sense, many transition phenomena affected by
temperature follow the Arrhenius equation. Examples include: <D transition of
electrons from a metal to a lightly-doped semiconductor Ca sub-atomic scale
transition); (2)diffusion of atoms Can atomic scale transition); @transition of a
Recent Developments in High-Temperature Permanent Magnet Materials 31

dislocation in crystals (a transition of an object with two dimensions in atomic


scale); and @transition of grain boundary (a transition of an object with one
dimension in atomic scale) . The domain wall transition is, in a sense, like the
transition of grain boundary. In both cases, the objects of the transition posses
only one dimension in atomic scale. However, when a domain wall moves,
atoms actually do not move. What moves is only a narrow region in which the
direction of magnetic moment differs from its neighbors. This is an important
difference between domain wall and other objects that follow the Arrhenius
equation. In addition, under a constant applied magnetic field, the thermal
energy would play a critical role in domain wall transition and the time-
dependent effect occurs. This would be the nature of magnetic viscosity
phenomenon.
(3) Explanation of Various Temperature Dependency of Coercivity Using
the New Model
According to this model, by modifying the compositions of the 1 : 5 cell
boundary phase and the 2 : 17 cell phase, various temperature dependencies
of the exponential /lE (T)/k B T can be obtained. There are six different
possibilities:
Q) With increasing temperature, /lE (T) decreases. In this case, /lE / k B T
decreases rapidly with temperature, which results in a large negative
temperature coefficient of M Hc. as observed in conventional 2 : 17 magnets
(Fig. 1.17, magnet 1);
CV With increasing temperature, /lE (T) increases at a smaller rate than
T.ln this case, /lE (T)/k B T decreases slowly with temperature, which
results in a small negative temperature coefficient of MH c ' as observed in some
newly- developed high-temperature magnets (Fig. 1.17, magnets 2, 3);
@ With increasing temperature, /lE (T) increases at the same rate as
T. In this case, /lE ( T) / k B T is virtually a constant and independent of T,
which results in a temperature-independent MH c ' as observed in some newly-
developed high-temperature magnets (Fig. 1.17, magnet 4);
@ With increasing temperature, /lE ( T) remains the same and is
independent of T, which results in a typical exponential relationship between
MH c and T;
@ With increasing temperature, /lE (T) increases at a larger rate than
T. In this case, /lE ( T) / k B T increases with temperature, which results in a
positive temperature coefficient of MH c ;
@ /lE( T) changes differently with temperature at different temperature
ranges, which results in a complex temperature dependence of coercivity as
observed in some newly-developed high-temperature magnets (Fig. 1. 17,
magnet 6).
As mentioned previously, coercivity mechanisms in permanent magnets
are very complicated and many factors affect the coercivity development. It is
believed that Eq. (1.7) reveals the primary nature of coercivity mechanisms in
2 : 17 type permanent magnets. Other factors that influence the coercivity,
32 Sam (Shiqiang) Liu

such as the geometry of the cellular structure and the width of the 1 : 5 cell
boundary phase, could be attributed to a modification of the constant c or the
addition of other terms in Eq. (1.7).
(4) Calculations of Temperature Dependence of MHc Using the Proposed Model
The temperature dependence of MH c was calculated using Eq. (1.7). In the
calculations, three liE( n cases were assumed. In each case, the calculated
result agrees with the observed experimental result.
Case I liE (n decreases with T. The assumed temperature depen-dence of
crystalline anisotropy constants of the 2 : 17 cell phase and the 1 : 5 cell
boundary phase is shown in Fig. 1. 38. The calculated result is shown in Fig. 1. 39.
This case results in a large negative temperature coefficient, as observed in
conventional 2 : 17 magnets (F ig. 1. 17, magnet 1).

300 400 500 600 700 800 900 1000 1100 1200
Temperature (K)

Figure 1. 38 Assumed temperature dependence of crystalline anisotropy constant for


2: 17 cell phase and 1: 5 cell boundary phase. Case I. i:1E(T) decreases with T.

300 400 500 600 700 800 900 1000 1100 1200
Temperature (K)

Figure 1. 39 Calculated temperature dependence of MHc Case I. i:1E (T) decreases with T.

Case n
liE ( n increases with T. The assumed temperature
dependence of crystalline anisotropy constants of the 2 : 17 cell phase and the
Recent Developments in High-Temperature Permanent Magnet Materials 33

1 : 5 cell boundary phase is shown in Fig. 1.40. The calculated result is shown in
Fig. 1.41. This case results ina small negative temperature coefficient, as observed
in some newly developed high-temperature 2 : 17 magnets (Fig. 1. 17, magnet 3).

~

u
&
80
"0 ==!
.~ ~

.5
~
u
c 1:5 cell boundary phase

300 400 500 600 700 800 900 1000 1100 1200
Temperature (K)

Figure 1. 40 Assumed temperature dependence of crystalline anisotropy constant for


2: 17 cell phase and 1 : 5 cell boundary phase. Case ]I, flEeT) increases with T.

300 400 500 600 700 800 900 1000 1100 1200
Temperature (K)

Figure 1. 41 Calculated temperature dependence of MHo. Case ]I ,


flE( T) increases with T.

Case ill /i.E (T) increases rapidly with T at high-temperature range


(500 to 800 K ). The assumed temperature dependence of crystalline
anisotropy constants of the 2 : 17 cell phase and the 1 : 5 cell boundary phase
is shown in Fig. 1.42, and the calculated result is shown in Fig. 1.43. This case
results in complex temperature dependence, as observed in some newly-
developed high-temperature 2 : 17 magnets.
The new model of coercivity can satisfactorily explain not only the
abnormal temperature dependence of MHo shown in Figs. 1.35 and 1.36, but
also the "normal" temperature dependence of MHo in conventional 2 : 17
magnets. This is because the two conflicting mechanisms exist not only in
34 Sam (Shiqiang) Liu

300 400 500 600 700 800 900 1000 1100 1200
Temperature (K)

Figure 1. 42 Assumed temperature dependence of crystalline anisotropy constant for


2 : 17 cell phase and 1 : 5 cell boundary phase. Case ill, I:>.E (n increases rapidly with T.

300 400 500 600 700 800 900 1000 1100


Temperature (K)

Figure 1. 43 Calculated temperature dependence of MH c . Case ill ,


I:>.E( n increases rapidly with T.

magnets with abnormal temperature dependence of MH c' but in all 2 : 17 type


permanent magnets. For conventional 2 : 17 magnets, the pinning mechanism
is relatively weak at high temperatures and the thermal activation mechanism
dominates at the entire temperature range. However, in the newly-developed
magnets, the pinning mechanism is so strong at high temperatures that the magnets
demonstrate small negative or even positive temperature coefficients of MHc .
It is believed that this thermal activation-pinning model also applies to
magnets that are controlled by crystalline anisotropy and domain wall pinning.
This concept may apply to all ferromagnetic materials in which domain wall
motion is involved during magnetization and demagnetization. For example, in
soft magnetic materials the coercivity is determined by the maximum
resistance to the domain wall motion. This resistance can be caused by any
crystal imperfection, including grain boundaries, inclusions, precipitates,
strain, and stress. The energy necessary for the domain walls to overcome an
Recent Developments in High-Temperature Permanent Magnet Materials 35

energy barrier llE (T) can be obtained not only from magnetic energy, but
also from thermal energy k B T. In general, the susceptibility of a soft magnetic
material may be expressed as
K cc exp[k B TIllE(T)]. (1. 8)

However, since the domain walls in soft magnetic materials are quite
thick resulting from the low magneto-crystalline anisotropy, thermal activation
of domain walls in soft magnetic materials is probably more difficult than in
magnetic materials with high crystalline anisotropy.

1. 4 New Approach to Calculate Temperature Coefficients


of Magnetic Properties

Permanent magnets with improved high-temperature performance are needed


for advanced systems. This requires not only developing new high-temperature
permanent magnets, but also a better and more accurate description of the
temperature characteristics of magnetic materials. The most commonly-used
parameter for temperature dependence of magnetic properties is the
temperature coefficient. The temperature coefficient of a magnetic parameter
Q over a temperature interval between T , and T2 is defined as

_ Q T2 - Q T, 0 I.
0"(T,-T2 ) - Q (T _ T) x 100 (~ C). (1. 9)
T, 2 ,

The temperature coefficient of Q at a specific temperature T is defined as

(1.10)

It is obvious that 0"(T , -T )


2
is an average of temperature coefficients of Q over
the temperature interval T, -- T 2 However, 0"(T,-T2 ) is not necessarily an
accurate description of the temperature dependence of Q, especially when the
interval between T, and T 2 is large. Further, when Q is not a monotonous
function of temperature T, Eq. (1. 9) may give a misleading result. On the
other hand, 0" T gives the temperature coefficient of Q at a specific
temperature T. It is the "true" (or instantaneous) temperature coefficient and
is a more accurate description of the temperature dependence of Q.
Unfortunately, in practice, it is impossible to calculate 0" T when II T a by =
simply using Eq. (1. 10).
The above-mentioned problem can be readily resolved if a polynomial is
used to represent Q,
n

Q( T) = 00 + 01 T + 02 T2 + ... + On Tn = ~ o/T/ (1.11)


/=0
36 Sam (Shiqiang) Liu

and redefine the temperature coefficient of 0 at T as

aT = ~~ x ~ x 100 (%/"C). (1. 12)

Coefficients 00' a" 02"'" an in Eq. (1. 11) can be determined using a least
squares fit. Calculating the derivative of this polynomial with respect to T yields

dT = a 1
dO + 20 2 T + 30 3 T 2 + ... + ~ io Ti-l
no n rn- 1 = LJ" (1. 13)
1=1

Substituting Eqs. (1.11) and (1.13) into Eq. (1.12) leads to


n

~ io,Ti-l
aT = -,-i=-''-n- - - X 100 (%/oC). (1. 14)
~O,T'
;=0

Using Eq. (1. 14), the "true" temperature coefficients a T of any magnetic
parameter 0 at any temperature T can be readily determined and a plot of
temperature coefficient versus temperature (a T vs. T) can be drawn.
Normally, since a T is more sensitive to T than 0, the aT vs. T plot is a very
useful tool to represent the temperature characteristics of a magnetic
parameter. The average temperature coefficient a (T, ~ T 2 ) can be also
calculated using Eq. (1. 11). The calculated results are very close to those
determined by conventional method.
As an example, Fig. 1. 44 shows the temperature dependence of mag-
netization at 10 kOe of a commercial sintered Gd2 (Co,Fe,Cu,Zr)17 magnet. In
the figure, the squares represent the experimental data, while the curve is a
6th degree polynomial fit. In any experimental characterization, random errors
are always associated with the results of measurements. The least square fit
eliminates those random errors and therefore the numerical result is generally a
better representation in comparison to the original experimental data.
Figures 1. 45 - 1. 47 are plots of temperature coefficients of magnetization,
intrinsic coercivity and maximum energy product versus temperature for some
8.0

7.5

8 7 .0
,:,!.

~ 6.5
o
.~
.t:! 6.0

"
~
c
~ 5.5
0: Experimental data
6th polynomial fit
5.0
4.5 '--_ _-'---_ _- L -_ _-----'-_-----'------'
o 200 400 600 800
Temperature (OC )

Figure 1.44 Temperature dependence of magnetization at 10 kOe of ~(Co,Fe,Cu,Zr)'7'


Recent Developments in High-Temperature Permanent Magnet Materials 37

sintered rare earth permanent magnets. This concept can be further developed
for modeling of temperature coefficients of magnetization for temperature
compensated rare earth permanent magnets. See Section 1.6 for further details.
0.4

Zero temperature coefficient


- -~--
1
2
1: Gd2 TM 17
2: Sm2TMl7
3: GdCo s
4: SmCo s
5: Nd-Fe-B
-1.0 ~--:-:..,--~~~=--~-e:-:-:--~"-----::-'
o 100 200 300 400 500
Temperature CC )
Figure 1. 45 Temperature coefficients of magnetization at 10 kOe of some
rare earth permanent magnets.

0.4
Zero temperature coefficient
---------.:::,.------

Figure 1. 46 Temperature coefficient of intrinsic coercivity of some


rare earth permanent magnets.

0.5
Zero temperature coefficient
Ei 0 ----------_:>..-_---
.~
Ei G -0.5
'0 ~ -10
'Ee....-
.~ g -1.5
S-g
~ is. -2.0
~ ~
3 ~ -2.5
I5 -3.0 Nd-Fe-B

~
r"
- 3.5 0L---:--!-:,..---:-~-~-=---:-'-:---':-:'-=---:-'.
100 200 300 400 600
Temperature CC )
Figure 1. 47 Temperature coefficient of maximum energy product of some
rare earth permanent magnets.
38 Sam (Shiqiang) Liu

1. 5 SmCo7 -Type Permanent Magnets

The Sm2Co17 compound has moderate high saturation magnetization and


magneto-crystalline anisotropy and the highest Curie temperature (920C)
among all rare earth-transition metal compounds. However, it has proved very
difficult to develop high coercivity in a pure Sm2 C0 17 compound. To develop
useful coercivity, a considerable amount of non-magnetic elements, such as Cu
and Zr, must be added, which results in significantly reduced magnetization
and energy product. Another disadvantage of the current Srnz (Co,Fe,Cu,Zr)'7
magnets is that a long-term heat treatment is necessary to develop high
coercivity.
SmC07 phase is a metastable binary phase in the Sm-Co binary system,
and it has a TbCu7 crystal structure which can be derived from SmCos(Wallace
and Narasimhan, 1980; Saito et al., 1989). The SmC07 compound has
saturation magnetization and a Curie temperature lower than a Sm2 C0 17 but
higher than a SmCos compound. It has been reported that the metastable
SmCO] phase can be obtained by melt spinning (Saito et al., 1989), splat
cooling (Buschow et aI., 1973), or mechanical alloying (Yang et al. , 1996).
It has also been reported that some transition metals, such as Zr and Ti can
stabilize S~ phase (Huang et aI., 1998). Recently, Zhou reported that high
coercivity and positive temperature coefficient of coercivity were obtained
without aging in SmCO]-x Ti x and SmC07- x- yCU X Tiycast alloys. If the Ti and/or
Ti, Cu substitutions can be made very low, and if the Fe substitution for Co can
effectively enhance the saturation magnetization as it does in the Srnz (Co, Fe) 17
system, then it would be possible for these 1 : 7-based compounds to have a
virtually higher saturation magnetization than Sm2(Co,Fe,Cu,Zr),7 magnets.
In addition, if high coercivity can be developed without long-term aging, it
would also be a very significant factor in for cutting the cost of production.
Therefore, efforts have been made to explore the possibility of synthesizing
sintered anisotropic magnet materials based on SmC07 phase (Liu et al. ,
2000d) .
It was found that a small amount of Ti addition stabilizes the 1 : 7 structure
in Sm(Col-v-x-yFevCuxTiy)z with v=0-0.15, x=0-0.14, y=O-O. 08,
and nominal z = 6. 1 - 7. O. Fe substitution for Co increases remanence, but
decreases coercivity. Cu increases coercivity, but decreases remanence. A
peak coercivity was obtained with x=0.105 in Sm(CObaICUxTio.043)64 (6.4 is
a nominal z value. Its corresponding effective z value is 7.0). The intrinsic
coercivity was found to be sensitive to Ti content. A peak coercivity of over 4
kOe was obtained when y = O. 04 in Sm (CObal CUo.oss Ti y )6.4 . On the other
hand, the magnetic properties were not very sensitive to the z value in
Recent Developments in High-Temperature Permanent Magnet Materials 39

SmCCoo.s72Cuo.oB5Tio.o43)z when nominal z = 6.1 - 7. O. The best room


temperature intrinsic coercivity obtained in a Cu-free magnet of
SmCCoO.957 Ti o.043 )6.4 was 1.8 kOe.lt was 4.5 kOe in a Cu-containing magnet of
SmCCooB75 CuO.OB5 Ti o.04 )6.4 and 7.1 kOe in a Hf modified SmCCoO.BO Cuo.16 Hfoo4 \.
The best room temperature maximum energy product CBH)max was 17.7 MGOe
CMegagauss Oersted) in a sintered anisotropic SmCCoo.soCuol6Hfo04)6
magnet. These properties were achieved after rapidly quenching from the
sintering temperature without aging. The coercivity mechanism remains
unknown. It is anticipated that further adjustments of composition and heat
treatment will lead to improved magnetic properties in sintered magnets based
on SmC07 phase.
As mentioned in Section 1. 3, the nominal z values of the newly
developed high-temperature magnets are also close to 7. The difference
between the aged SmTM_ 7 and the quenched SmCo_ 7-based magnets
described in this section is not only the fact that the latter may contain more
Sm. The fundamental difference is that the coercivity in the SmTM7 magnets is
developed in a long-period aging. These magnets have a fine-scaled cellular
microstructure consisting of three phases: a 2: 17 cell phase with a
rhombohedral structure, a 1 : 5 cell boundary phase with a hexagonal
structure, and a Zr-rich platelet phase with a hexagonal structure. While the
SmCo_ 7-based magnets are supposed to be single-phased alloys with a 1 : 7
hexagonal structure without any fine-scaled microstructure.

1. 6 Temperature Compensated Rare Earth Permanent


Magnets and Modeling of Temperature Coefficient
of Magnetization

Permanent magnets with very low-temperature coefficients of magnetic


properties over a wide temperature range are required for many applications.
Examples are microwave tubes, gyros, accelerometers, and conventional
moving-coil meters. The flux provided by most permanent magnets decreases
on heating. This is an intrinsic property for all ferromagnetic materials in which
the saturation magnetization will eventually drop to zero at Curie temperature.
To compensate for this magnetization loss with increasing temperature, partial
heavy rare earth substitution is used CBenz et a!. , 1974; Jones and Tokunaga
1976; Li et a!. , 1983; Liu, et a!. , 1990a, 1990b; Walmer et a!. , 1998). This
temperature compensation is based on the fact that in RE-TM compounds, the
magnetic moment coupling between RE and TM atoms depends on the type of
RE used in the alloy. When RE is a light rare earth such as Ce, Pr, Nd, Sm,
the moment couples parallel to the TM moment (ferromagnetic coupling). The
compound moment in this case exhibits a decrease with increasing temperature
40 Sam (Shiqiang) Liu

over the entire temperature range up to the Curie temperature. By contrast,


when the RE is a heavy rare earth such as Gd, Tb, Dy, Ho, Er, etc. , the
moment couples antiparallel to the TM-moment (ferrimagnetic coupling). For
ferrimagnetic materials, variations of the heavy rare earth (HRE )-TM
compounds with temperature are complex and it is possible that in a certain
temperature range the compound moment increases, rather than decreases,
with increasing temperature.
Figure. 1. 48 shows magnetization versus temperature (a) and temperature
coefficient of magnetization versus temperature (b), respectively. This figure defines
some important terms used in temperature compensation as follows:
T p : Temperature corresponding to the peak magnetization. Temperature
coefficient of magnetization (a) is zero at T p
T min to T max: Temperature range for optimum temperature compensation.
At T min the temperature coefficient of magnetization is {j, at T max it is - {j. The
magnitude of {j depends on appl ications. In most cases d < O. 005 % FC .
Good temperature compensated permanent magnets should have the
following characteristics:
(1) High T p or, for a particular application, a fixed T p
(2) Low T min' high T max or, large optimum compensation temperature
range. In this range, I a I ~ {j .
(3) Small {j.

Tp
T.min~_-r- T.
..:;m~ax~

Temperature
(a)

Zero coefficient
------~-
-5
I~Ptimum compensati~nI

Ternperature
(b)

Figure 1. 48 Temperature dependence of magnetization and


temperature coefficient of magnetization.

Gd has been used for temperature compensation in both 1 : 5 and 2 : 17


types of rare earth permanent magnets. Temperature compensated magnets
using partial Gd substitution for Sm illustrate very small temperature
Recent Developments in High-Temperature Permanent Magnet Materials 41

coefficients from - 50C to about 150C with a peak magnetization at around


room temperature. Higher T p and a large temperature range for compensation
are needed for many applications. In addition to Gd, other heavy rare earths,
such as Er and Ho can also be used for temperature compensation. Table 1.3
compares a few heavy rare earth 2 : 17 type compounds. It can be seen from
the table that Er2 TM 17 has the highest magnetization, while H02TM 17 has the
highest T p Previous experiments indicated that Gd2TM 17 has high coercivity,
but both Er2 TM 17 and H02TM 17 showed low coercivity. Thus, it would be
difficult to make a good temperature compensated magnet by using only a
single HRE. To obtain a temperature compensated permanent magnet with a
high coercivity, a high magnetization, a high temperature for the peak
magnetization and a large temperature range for compensation, it seems Gd,
Er, Ho, and probably more heavy rare earths would have to be used. This
typically requires considerable laboratory effort to determine the optimum
combination of the light rare earth (LRE) and the HREs. In research practice,
a method of blending powders is often used. For example, by melting only two
alloys of SmCos and GdCos ' any magnet alloys that have the composition of
(Sml- x Gd x ) Cos, with 0 < x < 1 can be obtained by blending powders of
SmCos and GdCos .

Table 1. 3 Comparison of a few heavy rare earth 2 : 17 type compounds.


Peak magnetization Temperature corresponding
Compound Coercivity
(emu/g) to peak magnetization, T p ( C )
Gel:! TM 17 60 300 High
Er2 TM 17 70 250 Low
H02 TM 17 63 400 Low

Because saturation magnetization (41TM s ) is an intrinsic property, it is


possible to calculate the temperature coefficient of 41TM s for a temperature
compensated RE-TM magnet using a simple model that the magnetization of an
(LRE 1 - x HRE x )-TM compound is independently contributed by LRE-TM and
HRE-TM. As a first step of the modeling, the temperature dependence of 41TM s
for SmCos and GdCos alloys should be experimentally determined by obtaining
two functions M 1 (T) and M 2 ( T). Then, two polynomials can be used to
represent these functions. Following that, these two polynomials can be "blended"
(added) instead of blending two alloys, obtaining a third polynomial,
(1. 15)
where 0 ~ x ~ 1. Next, the derivative of M 3 ( T) with respect to T,
dM 3 (T) j d T can be easily determined. Finally, the temperature coefficient of
the new "alloy" at any specific temperature can be derived using

aT
= dMdT(T)
3 x 1
M (T)
X 100 (OljC)
70.
(1.16)
3

In other words,the aT vs. T relationship for the new "alloy" can be readily established.
42 Sam (Shiqiang) Liu

The concept of calculating 4'ITM s in HRE-substituted rare earth permanent


magnets was first proposed by Camp et al (1985). and Ma et al (1986).
However, the approach they used for the calculation and the algorithms they
employed were not mentioned. S. Liu (1990a, 1990b) then proposed using a
polynomial to represent the temperature dependence of saturation
magnetization and moved one step forward by proposing to calculate the
temperature coefficients of (LRE 1- x HREx)-TM systems.
Using this approach, S. Liu et ai., (1999b) calculated the temperature
coefficients of magnetization of the (Sml- xGdx)C05 and (Sml- xGd x)2 (Co,Fe,
Cu, Zr) 17 systems with O<x< 1 in the temperature range of 20 to 600 C . In a
plot of aT vs. T, when aT changes its sign from positive to negative, the aT
vs. T curve crosses the line of aT = 0, and determines the temperature at
which the temperature coefficient becomes zero for magnets with various HRE
contents. The temperatures corresponding to the zero temperature coefficients
of the magnetization at 10 kOe applied field for (Sm'-xGd x)C05 and
(Sml- x Gdx ) 2(Co, Fe, Cu, Zr) 17 systems with 0 < x < 1 were determined and
the temperature compensation effect of Gd in these two systems was compared.
Figure 1. 49 is a plot of the temperature coefficient of magnetization at
10 kOe of (Sml- x Gd x) C05 versus temperature. The temperatures
corresponding to zero temperature coefficients for magnets with various Gd
contents are summarized in Table 1.4. The effects of Gd compensation on the
temperature coefficients of magnetization in (Sml-xGdx)Cos and (Sml- x Gd x)2 (Co,
Fe,Cu,Zr)17 systems are summarized in Fig. 1.50. Figure 1.50 also shows the
variation of magnetization of (Sml-xGdx)C05 and (Sml- x Gdx )2 (Co,Fe,Cu,
Zr) 17 versus Gd content. It is obvious from Table 1. 4 and Fig. 1. 50 that Gd
shows a much stronger effect for temperature compensation in the (Sm1 - x Gd x ) Cos
system than in the (Sml-xGdx)2 (Co,Fe,Cu,Zr)17 system. However, it must also
be noted that the temperature compensation effects are at the expense of
magnetization in both systems.
0.3
6
0.2

I: x=0.4 4: x=0.7
2: x=o.s 5: x=0.8
3: x=0.6 6: x=1.0

-{).3 '--_--'-_----'-_ _' - - _ - - ' - _ - - - - ' - _ . . - . J ' - - - - - '


o 100 200 300 400 500 600 700
Temperature ('C )

Figure 1. 49 Temperature coefficient of magnetization at 10 kOe of


(Sml- x Gd x ) Cos versus temperature.
Recent Developments in High-Temperature Permanent Magnet Materials 43

Thble 1. 4 Temperatures for zero temperature coefficient of magnetization for (Sml-. Gd. ) COs
and (Sml-.Gd.)2(Co,Fe,Cu,Zr)17 magnets.
Temperature for zero temperature coefficient ( C)
Gd content (x)

0.4 42 o
0.5 65 56
0.6 171 80
0.7 281 157
0.8 345 214
1.0 455 277

5 0 0 , . . . - - - - - - - - - - - - - - - - - , 12
f; 450
':::
c
400
.~ 350
6'
~ 300 ~
8 c
o 250 0
.~
~ 200 '.g
'0 6 c:

j., :~~ o ................... (Sml_.Gd.)Co s


5 ~~

E 50 ..... ----~ 4
.....................
~ 0 3
-50 '--_.L-_--'-_-'-_......L-_--'-_---'_----' 2
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Gd content, x

Figure 1. 50 Temperature for zero temperature coefficient of magnetization and


magnetization vs. Gd content in (Sml-.Gd. )C05 and (Sml-.Gd.)2 TM 17 systems.

The numerical expression of this approach is simple. Let us start from an


( LRE-HRE )-TM system. Suppose the temperature dependencies of
magnetization for LRE-TM and HRE-TM are represented by m-degree
polynomials:
m
M LRE-TM ( T) = ao + a, T + a2 T 2 + ... + am T m = ~ alT I = (T)
1=0
'I
(1.17)
m
M HRE-TM ( T) = b o + b 1 T + b 2T2 + ... + bmT m = ~blTI =
1=0
'2 (T).
(1. 18)
Coefficients ao, a" a2 , , am and b o ' b b 2 , , b m can be determined
using a least squares fit. Assuming that (n +" 1) experimental data points are
obtained for the LRE-TM magnet alloy and the temperature corresponding to
the j-th point is T I . Then, the magnetization corresponding to the j-th point is
44 Sam (Shiqiang) Liu

Y j (j = 0 to n). An m-degree least squares fit corresponds to solving the


following system of m + 1 linear equations:
n

(n + 1)00 + ~ T j0 1 + ~ n02 + ... + ~ Trom = ~ Yj


j=O
n

~ Tjo o + ~ nOl +~ n02 + ... + ~ Tr+ 1om = ~ YjT j


j=1 (1.19)
~ noo +~n Ol + ~ T1 0 2 + ... + ~ Tr+ 2o m = ~ Yjn

Since T j and Y j are obtained experimentally, 00' 01' 02' , Om can be


readily determined. Coefficients b 0' b l ' b 2 , , b m for the HRE-TM magnet
can be determined using the same procedure. Because, as mentioned before,
it is assumed that LRE- TM and HRE-TM independently contributes to the
magnetization of (LRE-HRE)- TM, the variation of magnetization versus
temperature for (LRE 1- x HRE x )-TM would be
M(LRE1-XHRExl-TM = (1 - x) 'I (T) + x '2 (T)
m m

= (1- x) ~O,T' + x~b,T'


1=0 1=0
m

= ~[(1 - X)OI + xb,JT ' = '3(T) (1. 20)


1=0

where x is the atomic fraction of the HRE.


The temperature coefficient of magnetization for the temperature-
compensated magnet is

= d['3(T)J x _1_ x 100(OLj'"')


aT dT '3(T) /0 '-' .

It is very easy to obtain the derivative of a polynomial,

d[f~~T)J= d~~[C1- X)OI + xb,JT '


m

= ~ i[ (1 - x) 0 1 + xb 1 J T 1-1 (1. 21)


1=1

Therefore,
m

~ i[(l - x)o/ + xb , JTI-1


aT = .:.c.1=-,--,~,-- X 100 (% Fc ) . (1. 22)
~[(1 - X)OI + xb/JT '
1=0

Now, let us consider RE-TM magnet alloys containing more than one HRE.
Suppose the target magnet alloy contains k (k >
2) rare earths, that is to consider

[(REI) Xl (RE2 ) X2 RE k )XkJ - TM


Recent Developments in High-Temperature Permanent Magnet Materials 45

where REI' RE z ' RE k stand for different rare earths (either LRE or HRE) and
Xl' Xz, "', Xk are their atomic fractions and Xl + Xz + ... + Xk = 1. The
variations of magnetization of REI - TM, RE z - TM, "', RE k - TM versus
temperature, determined by experiment, are represented by polynomials:
m

MRE,-TM = 4>, ( T) = ~QIITi


;=0
m

M RE2 - TM = 4>z ( T) = ~QZITI


1=0

M REk - TM = 4>k(T) = ~QkITI


1=0
Again, assuming the magnetization of [(REI) Xl (RE z) Xz'" RE K ) Xk ] - TM is
independently contributed by RE, - TM, RE z - TM, "', and RE k - TM within
the temperature range of the experiment, then the variation of magnetization
versus temperature for the magnet alloy is

M[(REI)XI(RE2lx2'''(REk)XkJ - TM
= XI4>I(T) + xz4>z(T) + ... + Xk4>k(T)
m

= ~(XIQlI + XZQZI + ... + XkQkl)T


1

1=0
m k

= ~(~XjQjI)T' = F(T). (1. 23)


1=0 j=l
The temperature coefficient of magnetization for the magnet alloy, aT' is

aT
= dFdT
(T) x _1_ x 100 (01
F(T) /0
je )
m k

~ i (~XjQjI )T
1 1
-
l
= I=lm j=k X 100 (%/"C). (1. 24)
~ (~XjQjl )T'
1=0 j=l

Very good agreement was obtained between the calculated data and the
experimentally determined results in (Sml- x Gdx)z (Co, Fe, Cu, Zr) 17 in the
temperature range of - 50 to - 300 'C. The soundness of this modeling
approach in more complicated systems is to be confirmed by experiments.
As mentioned before, temperature compensated magnets using partial Gd
substitution for Sm illustrate very small temperature coefficients from - 50 'C
to about 150 C with a peak magnetization at around room temperature.
Recently, in a temperature compensated magnet containing Er and Ho a very
high T p of 262 'C was obtained. This T p is more than 200C higher than that of
the (Smo.sGdo.s)z TM 17 magnet.
46 Sam (Shiqiang) Liu

1. 7 Other High-Temperature Permanent Magnet Materials


and Furture Perspective

In order to significantly improve the performance of permanent magnets, the


saturation magnetization (4nM s ) must be substantially increased since the
upper limit of the maximum energy product of a permanent magnet is
determined by (4nM s )2 /4. This would require an effort to find new compounds
containing more transition metal. Figure 1.51 illustrates saturation
magnetization versus the Co content in the binary Sm-Co system. The
saturation magnetization of LaC0 13 compound is also included in the figure. It is
obvious from Fig. 1. 51 that the saturation magnetization of the Sm-Co
compounds enhances linearly with increasing the Co content. As displayed in
Fig. 1. 51, the Curie temperature of the Sm-Co compound also enhances
linearly with increasing the Co content.
14

0: Experimental data
0: Estimated data
10.'-:-'<-------:':-_---:':-_---:':-_---:L-_-:-'
75 80 85 90 95 100
Co content (aL%)

Figure 1. 51 Saturation magnetization. 4lTM s of Sm-Co binary compounds vs.


Co content. 4lTM s of LaCo 13 is also given in the figure.

The LaC0 13 compound is the only stable 1 : 13 type compound found in all
binary rare earth-transition metal systems. Because it has high a concentration
of Co, the LaC0 13 compound possesses a very high Curie temperature of
1045 "C and a high saturation magnetization of 13 kG at room temperature.
Unfortunately, its cubic crystal structure and, hence, low magneto-crystalline
anisotropy prevented it from being developed into permanent magnets (Velge
and Buschow, 1968; Ido et aI., 1990). In recent years, Huang et al. (1995,
1996) studied the possibility of forming 1 : 13 type compounds in REC0 13 - x Si x
(RE = La, Pr, Nd, Gd, and Dy). Investigation should be carried out to
explore the possibility of forming a SmCo 13 compound by adding a third
Recent Developments in High-Temperature Permanent Magnet Materials 47

element and/or by using a special process because SmCo 13 may have the
highest magneto-crystalline anisotropy (see Section 1. 2). If successful,
another breakthrough in high-temperature permanent magnets is possible as
indicated in Table 1. 5. An assumption was made when estimating the
saturation magnetization of Sm(COO. 8 Feu)7 and Sm (CO O. 8 Feu) 13 compounds
that the effect of Fe on enhancing saturation magnetization in the 1 : 7 and
1 : 13 compounds is the same as that in the 2 : 17 compound. This was
partially confirmed in the 1 : 7 system as mentioned in Section 1. 5.
It will be not easy to develop the next generation of high-temperature
permanent magnets based on the SmCo 13 compounds. If SmCo 13 could be
successfully made, its magneto-crystalline anisotropy would be low, probably
much lower than those of Sm2 C0 17 and SmC05 compounds. It would then be a
challenge to increase the coercivity by, for example, substitution without
significantly decreasing the saturation magnetization and Curie temperature.

Table 1. 5 Saturation magnetization and Curie temperature of a few Sm-Co compounds.


4lTMs Curie point Theoretical
Compound Reference
(kG) Te("C) (BH)max
SmCos 11.4 727 32 (Strnat, 1988)
SmCop> 12.1 843 37 estimated
Sm(Ceo.8 Feu)7 (1) 13.7 801 47 estimated
Sm 2Co 17 12.5 920 39 (Strnat, 1988)
(3)
Sm2 (Ceo.8 Feo.2) 17 14.0 870 49
(2) (3)
Sm2 (Ceo.72 Feo2Cuo.osZro.02 h. 12.0 820 36
SmCo 13 (1) 13.3 1050 44 estimated
(1)
Sm(Ceo.8 Feu) 13 15.0 1000 56 estimated
(1) Metastable compound;
(2) Sintered magnet;
(3) University of Dayton magnetics Lab unpublished data.

References
Benz, M. G., R. P. Laforce, and D. L. Martin. In: Graham and J. J. Rhyne
eds. AlP Cont. Proc. No. 18/2: 1173(1974)
Buschow, K. , H. Broeder, and F. J. A den. J. Less-Common Met. 3: 191
(1973)
Camp, F. E., K .S. V. L. Narasimhan, and J. C.Hurt. IEEE Trans. Magn.
21: 1970(1985)
Chen, C. H. , M. S. , Walmer,M. H. , Walmer, S. Liu, and E. Kuhl. J. Appl.
Phys. 83: 6706(1998)
Fingers, R.T., and C.S. Rubertus. IEEE Trans. Magn. 36: 3373(2000)
Huang, M. Q. , W. E. Wallace, R. T. Obermyer, S. Simizu, and S. G. Sankar.
J. Magn, Magn. Mater. 151:150(1995)
Huang, M. Q. , W. E. Wallace,R. T. Obermyer, S. Simizu, M. McHenry, and
48 Sam (Shiqiang) Liu

S.G.Sankar. J. Appl. Phys. 79: 5949(1996)


Huang, M. Q., W. E. Wallace, and M. McHenry. J. Appl. Phys. 83: 6718
(1998)
Ido, H., J.C.Sohn. F. Pourarian. S.F.Cheng, and W.E.Waliace. J. Appl.
Phys. 67: 4978( 1990)
Jones, F.G., M.Tokunaga. IEEE Trans Magn. 12: 968(1976)
Khan, Y. Proc. 11th Rare Earth Res. Cont. n, 652 ( 1974)
Kim, A.S. J. Appl. Phys. 83: 6715(1998)
Lecrard, E. , C. H. Allibert, and R. Ballou. J. Appl. Phys. 75: 6277( 1994)
Li, D., J. Liu, S.Zhou, X.Jin and E.Xu. Proc. 7th Int'l Workshop on REPM:
495(1983)
Liu, J. F., T. Chui, D. Dimitrovv and G. C. Hadjipanayis. Appl. Phys. Lett.
73: 3007(1998)
Liu, J. F., Y. Zhang, D. Dimitrov and G. C. Hadjipanayis. J. Appl. Phys.
85: 2800 ( 1999a)
Liu, J.F., Y. Ding and G. C.Hadjipanayis. J. Appl. Phys. 85: 1670(1999b)
Liu, S., H.F.Mildrum and K.J.Strnat. J. Appl. Phys. 53: 2383(1982)
Liu, S., A.E. Ray and H.F.Mildrum. J. Appl. Phys. 67: 4975(1990a)
Liu, S., A. E. Ray and H. F. Mildrum. Proc. 11th Int'l Workshop on REPM,
389(1990b)
Liu, S. , G. P. Hoffman. IEEE Trans. Magn. 32: 5091( 1996)
Liu, S. , G. P. Hoffman and J. R. Brown. IEEE Trans Magn. 33: 3859 (1997)
Liu, S. , J., Yang G. Doyle, G. E. Kuhl, C. Chen, M. S. Walmer and M. H.
Walmer. IEEE Trans. Magn. 35: 3325 ( 1999a)
Liu, S., G.E.Kuhl. IEEE Magn. 35: 3271(1999b)
Liu, S., G Potts, G. Doyle, J.Yang, G.E.Kuhl, C.Chen, M.S.Walmer and
M. H. Walmer. IEEE Trans. Magn. 36: 3297(2000a)
Liu, S., J. Yang, G. Doyle, G. Potts, G. E. Kuhl, C. H. Chen,M. S. Walmer
M. H. Walmer. J. Appl. Phys. 87: 6728(2000b)
Liu, S., E. Kuhl, C. Chen, M. Walmer and W. Gong. Proc. 16th Int' I
Workshop of REPM, 159(2000c)
Liu, S., E. KUhl, C. Chen, M. Walmer and W. Gong. Proc. 16th tnt' I
Workshop of REPM: 71 (2000d)
Liu, S., G. Doyle, G. E. KUhl, C. Chen, M. Walmer and Y. Liu. to be
published in 2001
Livingston, J.D. and D.L.Martin. J. Appl. Phys. 48: 2608(1977)
Ma, B-M. , K. S. V. L. Narasimhan and J. C. Hurt. IEEE Trans. Magn. 22:
1081(1986)
Ma, B. M., Y. L. Liang, J. Patel, D. Scott, and C. O. Bounds. IEEE Trans.
Magn. 32: 4377(1996)
Mildrum, H.F., J. B. Krupar and A. E. Ray. J. Less-Commen Met. 93: 261
(1983)
Mishra, R. K. , G. Thomas, T. Yoneyama, A. Fukuno and T. Ojima. J. Appl.
Phys. 52: 2517(1981)
Recent Developments in High-Temperature Permanent Magnet Materials 49

Perkins, R. S. , S. Gaiffi and A. Menth. IEEE Trans. Magn. 11: 1431(1975)


Popov, A. G. , A. V. Korolev and N. N. Shchegoleva. Phys. Met. Metall. 69:
100(1990a)
Popov, A. G., V. S. Gaviko, L. M. Magat and G. V. Ivanova. Phys. Met
Metall. 70: 18 (1990b)
Ray, A. E. and K. J. Strnat. IEEE Trans. Magn. 8: 516( 1972)
Ray, A.E. J. Appl. Phys. 55: 2094(1984)
Ray, A.E., W.A.Soffa, J.R.Blachere and B.Zhang. IEEE Trans. Magn. 23:
2711(1987)
Ray, A.E. J. Appl. Phys. 67: 4972(1990)
Ray, A.E. and S.Liu. L. MatI. Eng. and Perf. 1: 183(1992)
Saito, H. , M. Takahashi, T. Wakiyama, G. Kodo and H. Nakagawa. J. Magn.
Magn. 82: 322(1989)
Schrefl, T. , J. Fidler, and W. Scholz. IEEE Trans. Magn. 36: 3394(2000)
Strnat, K. J. In: Wohlfarth, E. P. and K. H. J. Buchow. eds. Rare Earth-
Cobalt Permanent Magnets, Elsevier Science Publishers B. V. p.154(1988)
Velge, W.A.J.J. and K.H.J.Buschow. J Appl. Phys. 39: 1717 (1968)
Wallace, W. E. and K. S. Narasimhan. Science and Technology of Rare Earth
Materials, 395( 1980)
Walmer, M. S. , C. H. Chen, M. H. Walmer, S. Liu and G. E. Kuhl. Proc. 15th
Int'l Workshop on REPM, 689( 1998)
Walmer, M.S., C.H.Chen, M.H.Walmer, S.Liu, andG.E.Kuhl. Proc. 16th
Int'l Workshop of REPM, 41 (2000a)
Walmer, M. S. , C. H. Chen and M. H. Walmer. IEEE Trans. Magn. 36: 3376
(2000b)
Yang, J. , Mao, O. and Z. Altounian. IEEE Trans. Magn. 32: 4390 ( 1996)
Zhou, J. , I. A. AI-Omari, J. P. Liu and D. J. Sellmyer. J. Appl. Phys. 87: 5299
(2000)

Most recent advances in high-temperature permanent magnets related to the author and his
co-workers covered in this chapter were accomplished under the US Air Force and DARPA!
ONR sponsorship. Special gratitude goes to Dr. Christina Chen of Electron Energy
Corporation. Without her help and cooperation it would be impossible to finish this chapter.
The author would like to thank Dr. G. E. Kuhl of University of Dayton. Mr. W. Gong. and
Mr. M. Walmer of Electron Energy Corporation for their cooperation in the research
programs. The author's gratitude also goes to Ms. N. Maxwell of the University of Dayton for
editing the manuscript and to many of his colleagues who reviewed the manuscript and gave
valuable comments and suggestions.
2 New Rare-Earth Transition-Metal Intermetallic
Compounds and Metastable Phases for Permanent
Magnetic Materials

Z. D. Zhang

2. 1 Introduction

After three generations of the discovery of rare-earth permanent magnets,


based on SmCos ' Sm2 C0 17 and Nd2Fe14 B compounds, a great effort has been
made on researching new rare-earth transition-metal (R-T) intermetallic
compounds in permanent magnetic materials. The aim is to have a better
understanding of structural and magnetic properties of the intermetallics and to
find new R-T compounds possessing good intrinsic magnetic properties,
namely, a sufficiently high Curie temperature, high magnetization and high
uniaxial magnetic anisotropy.
Although it has been realized that there is a low probability for finding a
novel compound that would surpass Nd2Fe14 B in all of its favorable properties,
several advances has been made in this field, such as the introduction of
nitrogen and/or carbon into the lattice of R2 Fe17 or R (Fe, T) 12 compounds.
Scientists have been encouraged by realizing that the R2Fe14 B compounds are
not the only series of novel compounds to be found in the relatively unexplored
reservoir of ternary or multi-component compounds. Although there has been
some breakthrough on the R-T intermetallics, it seems that there is still a long
way to go before we reach the goal of finding the next generation of the rare-
earth permanent magnets. The difficulties are due to the following facts: First,
there exists the limitation for the formation of the stabilized R-T intermetallics.
Second, the intrinsic magnetic properties of one compound are constrained
due to the property of the elements that are necessary for stabilizing the
compound and the limitation of the symmetry of the crystal structure. Parallel
to the investigation of the intermetallic compounds, the study of new
metastable R-T phases has been an active direction in the area of the rare-
earth permanent magnets. New metastable R-T phases can provide more
chances for overcoming the second difficulty listed above.
A metastable phase does not exist in equilibrium conditions. The
metastable phase stabilizes under conditions corresponding to a local minimum
of the free energy. The large energy barrier between this local energy
New Rare-Earth Transition-Metal Intermetallic Compounds and ... 51

minimum and the lowest energy minimum hinders the formation of the
equilibrium phase. The study of the formation of metastable phases and the
transformation between metastable phases and the equilibrium phases is
helpful for understanding the relation between different crystal structures and
for understanding the thermodynamic stability of the intermetallics. Metastable
phases have been investigated extensively for the last two decades due to
their outstanding magnetic properties.
In this chapter, a comprehensive review of new R-T intermetallics and the
metastable phases in the rare-earth permanent-magnet materials is given. We
only focus on the R-T intermetall ics investigated recently, such as R (Fe, T,
X)12(T = V, Mo, etc; X = B, C), R2Fe14BNp R6Fe13-xM1+x(T = Fe, Co;
M = AI, Ga, Ge, Si, etc.) and multi-component compounds R3T29Si4Blo(T =
Fe, Co), and on the metastable R-T phases such as RFes' RT 7 (T = Fe, Co,
Ti, etc.) and their nitrides, R2Fe14C, R2Fe14BN6' RFe12 thin films. For a
detailed description of the progress in rare-earth permanent magnets, the
reader may refer to several review articles published previously (Buschow,
1977,1980,1988,1991,1997; Burzo, 1998; Coey, 1995; Franse and
Radwanski, 1993; Fuj ii and Sun, 1995; Herbst, 1991; Li and Coey, 1991;
Strnat, 1988). The reader may also refer to the author's recent review on the
metastable phases (Zhang et al. , 2000a, b). We do not focus on interstitial
compounds R3(Fe, T)29X and R(Fe, T)12X (X = N, C), which are represented
in other chapters of this volume. In the remainder of this section, we represent
an overview of the structures of metastable phases and related equilibrium
phases. The various processes employed for preparation of the intermetallics
and the metastable phases, such as arc-melting, mechanical alloying, rapid
quenching, a process of hydrogenation, disproportionation, desorption and
recombination (HDDR), solid state reaction, solid-gas reaction and
sputtering, will be introduced in Section 2. 2. The structural and magnetic
properties of the new R-T intermetallics, i. e., R (Fe, T, X)12 (X = B, C),
R2Fe 14 BN 6 , R6Fe13-xMl+x' and multi-component compounds R3T29 Si 4B IO , will
be presented in Section 2. 3. The formation of the metastable phases, the
transformation between them and the equilibrium phases, and the magnetic
properties of these phases will be shown in Section 2. 4. Sections 2. 5 and 2. 6
are for discussion and conclusion, respectively.

2. 1. 1 Overview of Structures

2. 1. 1. 1 Equilibrium Phases
The main phase of the first generation of the R-T permanent magnets is SmCos
compound, which has a family of isostructural compounds RCo s with a
hexagonal CaCus-type structure (the space group P6/mmm). In a unit cell,
the Ca (or R) atom occupies the 1a site, while Cu (or Co) atoms occupy 2c
52 Z. D. Zhang

and 3g sites. The CaCus-type structure is the basic structure of R-T


compounds, from which most of the structures of the R-T compounds can be
derived (Bouchet et al. ,1966; Buschow, 1971).
R2CO I7 , which is the base of the second generation of the R-T permanent
magnets, exhibits polymorphism (Khan, 1973). There are two modifications
of the structures: a high-temperature phase of the hexagonal Th2Ni l7 -type
structure (P6 3 / mmc space group) and a low-temperature phase of the
rhombohedral Th2Znl7 -type structure (R3m space group). There are the two R
sites (2d, 2b), the four T sites (4f, 6g, 12j and 12k) and the interstitial sites
6h and 121 in the unit cell of the hexagonal Th2Ni 17 -type structure. In the unit
cell of the rhombohedral Th2Zn17 -type structure, there exists the single R site
(6c), the four T sites (6c, 9d, 18f, and 18h) and the interstitial sites ge and
18g. The crystal structure of R2C0 17 of the Th2M 17 -type (M = Ni or Zn) could
be derived from that of RCos of the CaCus -type by an ordered substitution of a
dumbbell pair of two small Co-atoms in place of one large R atom. The
structure of the Th2 Ni 17 -type is built of blocks such as ABABAB, while the
Th2Zn17 is in order of ABCABC, where A, B, and C stand for substitution in
different rare-earth sites. R2 T 17 C x can be formed by arc melting, introducing
carbon atoms into the lattice of the R2T17 compounds (Zhong et a!., 1990). The
carbon atoms occupy interstitial sites in SI'l12 Fel7 Cx , which have a large effect on the
structure and magnetic properties of the compound.
The well-known R2Fe14 B has the tetragonal structure of Nd2Fe14 B-type
with space group P4dmnm. Each cell contains four formula units, 68 atoms.
There are six crystallographically distinct iron sites, two different rare-earth
sites and one boron site. For detailed description of the structure, the reader
is referred to the review of Herbst (1991). The structure of Nd2Fel4 B can be
viewed as stacks of triangular and hexagonal layers, which is analogous to the
hexagonal CaCus structure and a variety of other R-T phases. The hexagonal
arrays of Fe (k l ), Fe (k2), Fe (jl) and Fe (e) atoms in R2Fe,4 B are the
cognates of the TM (g) arrays in RTs ; both form hexagonal prisms enclosing
the R atoms. The B and Fe (h) sites of R2Fel4 B correspond to the TM (c) and
R sites in RTs,respectively.
The structure of RT I2 - x T x compounds is the tetragonal ThMnlrtype with
the space group 14/mmm (Florio et aI., 1952). The R atoms occupy the 2a
sites, whereas the T atoms lie in three sites 8f, 8i and 8j. The pure RFel2
phase does not exist in the equilibrium state and the tetragonal ThMnl2-type
structure can be stabilized by the elements of Ti, V, Mo, Cr, Mn, Si, AI,
etc. (Buschow et al. , 1988; '(ang et al., 1988; de Boer et al., 1987;
Verhoef et aI., 1988). The tetragonal ThMnl2-type structure is derived from
the hexagonal CaCus -type structure where half of the R atoms are replaced by
T dumbbells in an ordered fashion.
R3 (Fe, T)29 compounds with monoclinic Nd3 (Fe, Ti)29-type structure have
been characterized as belonging to the space group A2/m' a mixed structure of
1 : 12 and 2 : 17 phases (Shcherbakov et aI., 1992; Hu et aI., 1994;
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 53

Kalogirou et al . 1995; Pan et al.. 1998). This is a minimal non-isomorphic


supergroup of P2 1/c (Kalogirou et al.. 1995). This family of R-T compounds
is derived from the hexagonal CaCus-type structure by replacing two fifths of R
atoms by the T dumbbells.
The presence of the R6T 13 - x M1+x compounds (T = Fe. Co; M = Cu. Ag.
Au. AI. Ga. Ge. Si and Sn) as a secondary phase enhances the coercivity of
Nd-Fe-B permanent magnets. The R6 Fe13-x M1+ x compounds were found to
order in a tetragonal L~COll Gartype structure with space group 14/mcm
(Sichevich et al.. 1985). The rare earths occupy two crystallographic sites.
8b and 161. while the two other elements together occupy five sites. 4a. 4d.
16k. 161, and 1612 in the La6 COll Ga3 -type structure. Nd6Fe'3 Si crystallizes as
an ordered variant of this type of compounds (Allemand et al. 1990).
A series of multi-component compound R3T29 Si 4B lO (T = Co. Ni) was
discovered (Wu et al.. 1993; Zhang Heng et al.. 1997). which establish a
link between the ternary silicides and borides. The compound has a tetragonal
Nd3Ni29Si4BlO-type structure with space group P4/nmm (Bulcock et al.. 1998;
Zhang Heng et al.. 1998. 1999), which does not form for T = Fe (Zhang
Heng et al.. 1999).
2. 1. 1. 2 Metastable Phases
Light rare-earth metals do not form compounds of CaCus-type with iron in the
equilibrium states. The Th2Zn17-type rhombohedral structure exists in samples
of nominal composition SmFes and SmFe7 (Ray. 1966). The metastable
SmFes and NdFes phases have been found to form in the alloys prepared by
rapid solidification. which have the hexagonal CaCus-type structure with the
space group P6/mmm. isostructure of the RCos compounds.
The structure of R2T 17 reduces to that of the TbCu7-type (P6/mmm space
group). when a dumbbell of T atoms replaces in a disordered fashion some of
the R atoms in the CaCus-type structure. The' TbCu7-type RC07 phase is a
high-temperature phase stabilized at temperatures above 1300 "C for several
rare-earth elements. At lower temperatures. ordering sets in and the high-
temperature disordered phase transforms into R2C0 17 of the Th2 Ni 17 -type or
Th2Zn17-type structures. The Th2M 17 -type structure (M = Ni or Zn) of R2C0 17
is superstructure of the TbCu7 -type structures of the disordered RC07 phases.
The structural and magnetic properties of RT7 (R = Fe. Co) compounds
depend sensitively on the disordered replacement of the dumbbell of T atoms.
R2Fe14 C compounds do not crystallize from melting and must be formed by
a peritectoid-Iike solid-solid transformation during annealing ingots or melt-
spun precursors or mechanically alloyed powders. The R2Fe 14 C compounds.
isostructural with R2Fe 14B-type structure. could be catalogued to the kind of
metastable phases.
Introducing nitrogen or carbon atoms into the interstitial sites of the
R2T 17 and RT,2-x T x lattices leads to the formation of R2T,7Ax and RT 12 - xTxA y
(A = N. C) (Coey and Sun, 1990; Yang et al. ,1991a, 1991b; Coey et al.
54 Z. D. Zhang

1991; Kou et al. , 1991 ). The A atoms occupy the ge and/or 18g (6h and/or
120 sites of the Th2Zn17-type CTh2Ni 17 -type) R2T 17 lattice up to a maximum of
six per formula unit, while the A element fills the 2b site in the ThMn l2-type
structure to a maximum of one per formula unit. The octahedral 2b interstitial
sites in the ThMn12-type structure are equivalent to the ge or 6h sites in the
Th2Znl7 -type (or Th2Ni l7 -type) structure.
The pure RFel2 phase does not exist in the equilibrium state and the
tetragonal ThMnl2-type structure is stabilized by adding the elements Ti, V,
Mo, Cr, Mn, etc. DC magnetron sputtering was used to fabricate SmFe'2 films
with ThMnl2-type of tetragonal structure (Cadieu et aI., 1991; Wang et al. ,
1993). In this case, the ThMn12-type structure in the SmFel2 films is metastable.

2.2 Experimenfal Processing

The equilibrium phases such as intermetallic compounds can be prepared by


cooling the liquid of alloys with a very low rate of cooling. An alternate
approach for reaching the equilibrium phases is to anneal the as-cast ingots at
elevated temperatures for time as long as possible. Arc-melting or induction-
melting and subsequent annealing are the convenient processes for the
preparation of the intermetallic compounds.
Various processes have been used to prepare metastable phases. The
most convenient method is first to form the amorphous phase and then to
anneal it at an appropriate temperature. The metastable phases crystallize
normally at temperatures slightly higher than the crystallization temperature. It
is necessary to choose the annealing temperature to be as close to the
crystallization temperature as possible, in order to promote the formation of
most metastable phases. Usually, annealing at higher temperatures leads to
either the growth of the crystal grains or the transformation to more stable phases.
The following processes are those based on these methods: Q)mechanical
alloying; (2) mechanical milling; @ rapid quenching; and @ the process of
HDDR. Solid-state reaction and solid-gas reaction are other two powerful
processes for forming the metastable phases. DC magnetron sputtering was
used to fabricate metastable SmFe12 films with ThMnl2 -type of tetragonal
structure (Cadieu et al. ,1991; Wang et al. ,1993) .In this section, we briefly
introduce these processes.

2. 2. 1 Melting
The process of melting is a tradition method for preparation of the alloys and
the intermetallics compounds. Starting materials with high purity are melted by arc-
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 55

melting or induction melting under purified argon. The ingots are then annealed at a
certain temperature for a long time for reaching the equilibrium phases.

2. 2. 2 Mechanical Alloying/Milling
Mechanical alloying uses mechanical energy to activate chemical reactions and
structural changes to form alloys directly from elemental precursors
(Benjamin, 1970, 1976). The repeated collisions between balls and powders
with very high impact velocity in mills cause the powder particles to be
deformed and work-hardened. Cold-welding of overlapping particles occurs
between atomistically clean surfaces formed by prior fracture. The
continuously competing processes of deformation, fracture and welding change
composition and refine microstructure. Mechanical alloying has been
extensively applied to synthesis rare-earth permanent magnets (Schultz et al. ,
1987; Schultz and Wecker, 1988; McCormick, 1997; McCormick et aI.,
1996). Mechanical alloying is usually carried out in an inert atmosphere to
prevent the rare-earth alloys from oxidation. Mechanical alloying can
overcome the diffusion limitation that prevents most solid-state reactions from
occurring at low temperatures. The amorphous phase may occur after
mechanical alloying for a milling time long enough in a certain strength of
milling powers. The subsequent annealing after mechanical alloying is
beneficial to the formation of metastable phases at relevant temperatures. A
similar process is mechanical milling, which starts from the already alloyed
materials (Zhang et al. , 2000a,2000b).

2. 2. 3 Rapid Quenching
Rapid quenching has been employed frequently to synthesize the amorphous,
nanocrystalline and metastable phases. The liquid alloys are quenched rapidly
on the surface of the water-cooled copper rollers. The composition of the
alloys and the velocity of the rollers control the formation of the amorphous,
nanocrystalline, and metastable phases. Normally, for a certain composition,
the faster the velocity of the rollers is, the smaller the grain size. If the cooling
rate is fast enough, the amorphous phase could form directly after the rapid
quenching. Metastable phases may form after annealing the amorphous ribbons
at temperatures higher than the crystallization temperature or even exist
already in the as-quenched ribbons.

2.2.4 HDDR
The process of HDDR has been used to form the amorphous, nanocrystalline,
and metastable phases. Hydrogenation at temperatures higher than about
56 Z. D. Zhang

550 C leads to the disproportional reaction of hydrogen with rare-earth


intermetallics to form rare-earth hydride and a-Fe phase. The subsequent
desorption and recombination processes lead to the formation of nanoscale
isotropic rare-earth permanent magnets. The phases formed in the HDDR
process are usually metastable, which could transform into the equilibrium
ones after anneal ing at higher temperatures.

2. 2. 5 Solid-State Reaction
The solid-state transformation from the high-temperature phase to the low-
temperature phase offers the possibility of generating fine grain
microstructures in some cast alloys. The advantage of this technology is its
simple production route. For instance, the R2Fel4 C phases with light rare
earths are normally not found in as-cast alloys and are formed at lower
temperatures during annealing by a peritectoid-Iike transformation (Liu et al. ,
1987). The decomposition temperature at which the R2 Fel4 C compounds
transform by solid-state reaction into primarily R2Fel7 C x increases as a
function of the atomic number of the R component (Eisses et al., 1991). The
solid-state transformation after mechanical alloying extends the composition/
temperature range for the formation of light rare-earth R2Fel4 C phases (Sui
et ai., 1996, 1997b, 1997d).

2. 2. 6 Solid-Gas Reaction
Solid-gas reaction is usually carried out at a temperature that is suitable for the
reaction of the gas and the powders of the alloys. The velocity of reaction is
controlled by the temperature, the pressure of gas, the particle size of the
alloy, the activity of the surface/interface of the powders, the diffusion rate of
the gas in the alloy, the number of the possible interstitial sites in the lattice,
etc. Solid-gas reactions, such as nitrogenation and carbonation, are very
powerful for synthesizing new rare-earth permanent magnets. The most
successful example is the discovery of rare-earth-iron-based intermetallic
compounds Sm2 Fe17 N3 with high-performance magnetic properties (Coey,
1995). However, the rare-earth-transition-metal nitrides (or carbides) are
highly unstable, which decompose into the rare-earth nitrides (or carbides)
and a-Fe phase at temperatures of 600 "C and above.

2. 2. 7 Sputtering
Sputtering is a convenient method for preparing films. Several SmCo-based
films with various structures were synthesized by sputtering (Chen et al. ,
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 57

1993; Hegde et al., 1996). Magnetic hardening in SmCo x-Co and PrCo: Co
multilayers and nanocomposites were studied CUu et aI., 1999). RFe12 films
have been synthesized by sputtering CCadieu et al., 1991; Wang et al. ,
1993), which are metastable since the ThMnrtype RFe12 phase does not exist
in its equilibrium state.

2. 3 New Intermetallic Compounds

Searching for new intermetallic compounds is an effective way for making


better permanent magnets. The discovery of Nd2Fe14 8 is a well-known
example for this purpose. However, through the long term effort, scientists
worldwide realize that it is quite difficult to find a novel compound that would
surpass Nd2Fe14 8 in all of its favorable properties, Nevertheless, the study of
new intermetallic compounds is still important for better understanding of
structural and magnetic properties of the intermetallics. Scientists have been
encouraged by realizing that there are a lot of unknowns in the relatively
unexplored reservoir of ternary and multi-component compounds.
In this section, we focus on the R-T intermetallics investigated recently,
such as RCFe,T,X) 12 CX = 8, C), R2Fe148Ns' R6Fe13-xMl+x' and multi-
component compounds R3T29 Si48 10

2.3.1 R(Fe,T,X)12(X = B, C)
The effects of boron and carbon on the structure and magnetic properties of
ThMn'2-type rare-earth transition-metal intermetallics were studied by Chuang
et al. C1995), Zhang et al. C1995) and Dong et al. C1996). It was found that
carbon atoms occupy not only the interstitial 2b site, but also the substitutional
8i site in the ThMn,rtype lattice. The lattice parameters a and c of
YCFeVC)12 are slightly increased, and subsequently the unit cell volume
expansion is 1.4 % compared to the carbon free YFelO V2 . This volume
increment of 1. 4 % for the arc-melted carbide is slightly lower than that of 1. 7 %
reported for YFe,oV2Cx produced by gas-solid reaction using ethane CC2H6 )
and/or methane CCH 4). Consequently, the increase in the Curie temperature
is lower in the arc-melted carbide. The lower volume expansion and the
smaller increase of the Curie temperature originate from the lower occupation
at the interstitial 2b site. On the other hand, it was observed that the boron
atoms occupy the 8i sites, substituting for a few Ti atoms, while the
substituted Ti atoms enter the interstitial 2b sites in the tetragonal ThMnlrtype
structure.
58 Z. D. Zhang

The Curie temperature and magnetic properties at 4.2 K of RCFeVB) 12 ,


RCFeVC)12 and RFelO V2 CR = Y, Nd or Sm) compounds are listed in
Table 2. 1. The Curie temperatures of RCFeVB)12 compounds CR = Nd, Sm,
Y) are enhanced compared with the boron free compounds. In some of these
compounds the saturation magnetization M s at 4.2 K is also increased. A spin
reorientation occurs at 130 K in Nd CFeVB) 12. Compared to that C150 K) of
NdFelO V2 , the decrease of the spin reorientation temperature indicates that the
planar anisotropy of the Nd sublattice decreases with the B addition. First
order magnetization process CFOMP) transitions occur at 4. 2 K in both
SmCFeVB) 12 and Nd CFeVB) 12 compounds, with onset fields of 11. 5 and
2.2 T, respectively. The anisotropy fields of the borides are the highest
among these compounds, which are 20, 4 and 5 T for Sm CFeVB ) 12 ,
NdCFeVB)12 and YCFeVB),2' respectively. The enhancement of the anisotropy
field found for the borides may be caused by occupation of the boron atoms at
the substitutional 8i site.

Table2.1 Curie temperature and magnetic properties at 4.2 K of R(FeVB)12,R(FeVC)12


and RFe 10 V2(R=Y,Nd or Sm)compounds.
Compounds Tc(K) M. (/.Ia/f. u.) /.IR(1) (/.Ia) /.IF. (/.Ia) /.1M (/.Ia) BA(T) B FOMP (T)

Nd(FeVB)12 607 22.9 3.3 1. 96 1. 64 4 2.2


NdFe,o V2 577 18.5 3.3 1. 52 1. 26 3 2.3
Sm(FeVB)'2 632 19.2 0.7 1. 85 1. 54 20 11.5
Sm(FeVC)'2 612 19.3 0.7 1. 86 1. 54 18 -8
SmFelO V2 603 19.2 0.7 1. 85 1. 54 14 11. 0
Y(FeVB)12 568 16.9 0 1. 69 1.4 5
Y(FeVC)12 547 16.4 0 1.64 1. 37 4.5
YFel0 V2 534 15.9 0 1. 59 1.3 4
(1) IIR is the magnetic moment of rare-earth ion, and 11M the magnetic moment of transition-metal
atom.

The anisotropy fields of SmCFeVC),2 and YCFeVC)12 are 18 and 4.5 T,


which are slightly higher than those of the corresponding carbon-free
compounds SmFel0 V2 and YFelO V2. This increase may be ascribed to the
contribution of the carbon atoms in the 8i site. It implies that the effect of the
substitutional carbon at the 8i site on the crystalline electronic field overrides
the contribution of the interstitial carbon at the 2b site, resulting in an increase
in the anisotropy field. The values for the magnetization at 4.2 K of
SmCFeVC),2 and YCFeVC) 12 are 19.25 PB/f.U. and 16.42 PB/f.U., increase
insignificantly by 0.05 PB/f.u. and 0.52 PB/f.U., respectively, for SmFe ,O V2
and YFelO V2 , owing to the introduction of carbon atoms. The magnetization
curve in the hard direction at 4.2 K of SmCFeVC),2 shows a FOMP-Iike or
quasi-FOMP-like transition.
New Rare-Earth Transition-Metal Intermetallic Compounds and... 59

Since R2Fe14B was discovered (Sagawa et al. , 1984), only has carbon been
reported capable of replacing boron in R2Fe14 B and forming carbides with the
Nd2Fe,4B structure (de Boer et ai., 1988). The nitrides R2Fe14BNx(R = y,
Nd) were prepared by gas reaction (Zhang et ai., 1993). Quaternary rare-
earth transition-metal boron nitrides R2Fe'4 BNo. 1 (R = Nd and Sm) were
synthesized by arc melting (Zhang et al. 1999). The structure and magnetic
properties of these quaternary compounds have been studied by means of
X-ray diffraction, neutron diffraction, ac initial-susceptibility and magnetization
measurements. The bond of boron nitride (BN) can be broken by arc-melting
to combine with rare-earth and transition-metal atoms. Most nitrogen atoms
(about 90 % of the starting nitrogen) vapor out of the samples during arc-
melting. Only about 10% nitrogen remains in the compound after arc-melting,
which may occupy the interstitial site (40, as prepared by gas reaction
(Zhang et al. , 1993; Yang et ai., 1991c), or absorb in the grain boundaries.
The R2Fe14 BNo. 1 (R = Nd and Sm) compounds have the tetragonal Nd2Fe14 B -
type structure with space group P4 2/ mnm.
Results of the magnetization measurements, at 1. 5 and 293 K, of the
Nd2 Fe14 BNo. 1 compound are shown in Fig. 2. 1. The room temperature anisotropy
fields of the boron nitrides are slightly larger than those of Nd2Fe14 B (Sagawa
et ai., 1984; Herbst, 1991; Burzo, 1998; Liu et ai., 1990). Lattice parameters
(derived from X-ray diffraction) , Curie temperatures, saturation magnetization,
and anisotropy field of the boron nitrides are listed in Table 2.2. From Table 2.2,
the addition of nitrogen atoms does not obviously expand the lattice of the
180

o 234 5 6 7
B(T)

Figure 2. 1 Magnetization curves with magnetic field direction perpendicular ( x) and


parallel ( ) to the aligned direction for N~ Feu BNo. 1 compound at 1. 5 K (solid lines) and
293 K (dotted lines) (After Zhang et al. ,1999).
60 Z. D. Zhang

Nd2Fe'4 B compound, due to the fact that only few nitrogen atoms remain in the
samples. Consequently, the effect on the increase of Curie temperature by
such addition is not very evidently compared with that by gas reaction (Zhang
et al. , 1993). Curie temperatures and anisotropy fields at room temperature of
the R2Fe'4 B compounds are slightly increased by the incorporation of nitrogen.
The addition of the nitrogen atoms gives a weak negative effect on saturation
magnetization at 4.2 K of Nd2Fe'4 B. The negative effect on the saturation
magnetization becomes more pronounced at 293 K. Temperature dependence
of magnetization, measured at 6. 5 T, of Nd2Fe'4 BNo.05' Nd2 Fe'4 BNo. 1 ,
Sm2Fe,4BNo., is represented in Fig. 2.2, where temperature dependence of
magnetization, measured at 0.03 T, of N~ Fe'4 BNo.1 is presented as the inset. The
inset shows that in Nd.!Fe'4BNo.1 an SR transition occurs at 135 K that is the same as
that of Nd.!Fe'4B (Herbst, 1991; Burzo, 1998). The SR is also found to occur at the
same temperature in N~ Fe'4 BNo. 05 .

Table 2. 2 Lattice parameters (derived from X-ray diffraction), Curie temperatures,


saturation magnetization, and anisotropy field of the R2Fe,. BN x compounds with A = Nd and
Sm (After Zhang et al. , 1999). The two columns for anisotropy fields were derived from the
intersection points of the extrapolated magnetization curves, and the Singularity Point
Detection SPD curves, respectively.
Lattice parameters M . 1.5 K
Compounds Tc(K) BA.293K (T)
a(A) c(A) (A rn2 /kg) (lJe/ f . u.)

N~Fe,.B 8.79 12.15 586 190 37.7 7.3 8.1


N~ Fe,. BN o.05 8.78 12.18 590 180 35.0 10.7 8.8
N~Fe,.BNol 8.79 12.20 596 175 34.4 10.3 8.8
Sm2 Fe,.B 8.80 12.15 616 168 33.3 15.0 -
Sm2 Fe,. BN o., 8.79 12.20 584 168 33.3 10.5 8.5

180

'#E 160
:':::

~,
-
~ 140
r~~2iO::~~~~~~l:l:l:l:::::::::i'Jlo'"
:l19
'"
E 18

:s~ 1716 .
Nd2Fe\4 BNo.\
,

.......
.....
15 '---'----'---'-----'--
o 50 100 150 200
T(K)
120 oL-----...,.1~00=------2:-:0'-::0----~300
T(K)
Figure 2.2 Temperature dependence of magnetization measured at 6.5 T of
Nd2Fe,. BNo.05 ( ) , Nd2Fe,. BNo., (+) and Sm2 Fe,. BNo., (X) compounds. The inset
shows temperature dependence of magnetization measured at O. 03 T of N~ Fe,. BNo. 1
compound (After Zhang et al., 1999).
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 61

The crystal structure of R6Fe13- x M 1+x compounds (T = Fe, Co; M = Cu, Ag,
Au, AI, Ga, Ge, Si, and Sn) was unraveled, with orders in a tetragonal
La6 COli Ga3 -type structure with space group 14/ mcm (Sichevich et al. , 1985).
Nd6Fel3 Si crystallizes as an ordered variant of this type of compound
(Allemand et al. , 1990). The R6FeI3-xMI+x compounds are interesting from a
technological point of view because their presence as a second phase
enhances the coercivity of Nd-Fe-B permanent magnets.
Magnetic properties of the R6 Fe13- x M l +x compounds and their hydrides
were investigated (de Groot et al., 1998). A large hysteresis in the field
dependence of the magnetization is present in all compounds including
La6 Fell Ab, indicating the role of the Fe-sublattice anisotropy. Figure 2.3 gives
field dependence of the magnetic moment at 4. 2 K of the free powder of
L~ Fell A1 3. Crystal structure, magnetism and 57 Fe Mossbauer spectra of
~ Fel3 M compounds were studied (Hu et aI., 1992). Hydrogen induced metamag-
netism and the change in magnetic structure with hydrogen in R6Fe13 M
compounds were attributed to a change of sign of the weak antiferromagnetic
coupling of the iron sublattice (Coey et aI., 1994). Magnetic properties of
R6Fel3Sn (R = Nd, Pr) compounds were studied by Xiao et al. (1998).
20
18
16
14

Figure 2. 3 Field dependence of the magnetic moment at 4. 2 K of the free powder of


L8G Fell A1 3 The line corresponds with data taking during continuous sweeps with
increasing and decreasing field (After de Groot et ai., 1998).

The search for new multi-component intermetallics is one of the active


directions in the field of the rare-earth permanent magnets. Upon the long-term
62 Z. D. Zhang

efforts of the scientists in this field, most binary rare-earth-transition-metal


systems have been explored and most binary compounds existing in the binary
systems are well-known. It is very difficult to find novel binary intermetallics
with high magnetic properties in these well-known systems. On the other
hand, it has been bel ieved that the Rz Fe,4 8 compounds are not the only series
of novel compounds with high performance in the relatively unexplored
reservoir of ternary and multi-component compounds. A great interest has
been attracted to the synthesis of novel multi-component compounds. This
effort is beneficial, not only to the synthesis of the new compounds, but also to
the understanding of the structural and magnetic behaviors of the elements in
the multi-component systems.
Recently, a novel series of R3 TZ9 Si 48 10 (T = Co, Nj) compounds have
been discovered (Wu et al., 1993; Zhang Heng et al., 1997), which
establish a link between the ternary silicides and borides. These compounds
have a tetragonal Nd3 Niz9 Si 48 1O -type structure with space group P4/nmm
(8ulcock et aI., 1998; Zhang et aI., 1999). It has been found that the
compound does not form for T = Fe (Zhang et aI., 1999). X-ray diffraction
and SEM observation confirmed the single-phase character for the samples
investigated (Zhang Heng et al., 2000a). Figure 2. 4 shows the lattice
parameters a and c of R3 CO Z9 Si 48 10 compounds (Zhang Heng et al. , 2000a).
With the continuous shrinking of the radius due to the filling up of the 4f shell,
the lattice parameters for the compounds exhibit a monotonic decrease with
increasing the atomic number for the rare-earths except for Ceo An anomaly for
the lattice parameters against atomic number was observed for the ~ Co.z9 Si 48 10
compound, similar to Ce-based compound in other systems, which was
ascribed to the mixture of the + 3 and + 4 states of Ce ion.
~ 11.31
~
(l) ---<II-- a
11.26
~
0..
8 1I.21
.~

....l 11.16 L..L--'-_'----'---'---':c-----"L-i.----'-------'------'-----''-


R
(a)

j
i8
.~

....l 7.83 '-:-L--::'----:!-:-!--c---::'--::!-:-=':-::!-:7--=----=-----=-L-


La Ce Pr Nd Sm Gd Tb Dy Ho Er Tm Lu
R
(b)

Figure 2. 4 Lattice parameters a and c of R3 C0z9 Si. 8'0 compounds


(After Zhang Heng et al. , 2000a).
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 63

Some compounds exhibit paramagnetism at room temperature.


Nd3 C029 Si 4 8 10 exhibits a magnetic transition around 200 K with a second
magnetic transition occurring around 43 K (Campbell et aI., 1998). Addition
of -- O. 5 wt. % 57 Fe decreases the magnetic ordering temperatures to about
167 and 38 K, respectively (Campbell et aI., 1998; Wu et aI., 1998). The
former is likely due to its Curie temperature, while the latter is likely due to a
spin reorientation as a result of the temperature-induced competition between
the rare-earth sublattice and transition-metal sublattice anisotropies of different
preferential directions, or between different order crystal fields in terms of the
rare-earth sublattice (Campbell et al. , 1998).
A Mossbauer study on the 57Fe doped R3 T29 Si 4 8 1O compounds
demonstrated that 57 Fe atoms preferentially occupy the 2c crystallographic
site; the 8jl, 8j2 and 8i2 sites are the second preferentially occupied and the
8i3 and 16k sites are the third preferential occupations (Wu et aI., 1998;
Zhang Heng et al. , 2000b). The magnetic hyperfine fields derived from the fit
of the spectra at 4. 2 K are shown in Fig. 2. 5, and they increase with the
number of unpaired 4f electrons (Zhang Heng et al. , 2000b). The variation of
the magnetic hyperfine fields versus the atomic number of the rare earths was
attributed mainly to the difference of the R-Co interaction, after removing the
size effect by a linear subtraction as shown in the dashed line in Fig.2.5.

28
b
j
27

26 '-;-L----;:!---;!--::+.-+~;--;:;:;__;~:-';-__::!___;;:!_-----f
La Ce Pr Nd Sm Gd Tb Dy Ho Er Tm Lu
R3C029Si4BIO compounds

Figure 2. 5 Magnetic hyperfine fields derived from the fit of the MOssbauer spectra at 4. 2 K
of R3 GO(57 Fe)29 Si 4 8 10 compounds. The gap between the solid line and dashed line
represents the effect of unit cell size on the magnetic hyperfine fields due to the lanthanide
contraction (After Zhang Heng et al. , 2000b).

Although there has been great progress in researching/studying the multi-


component compounds, it is still too far from their possible application for the
purpose of permanent magnets, because of their comparatively low Curie
temperature, magnetization, and anisotropy. It is expected that long-term
efforts are still needed before the realization of the goal.
64 Z. D. Zhang

2. 4 Metastable Phases

It has been become clear that studying the metastable R-T phases will be one
of the most active directions of permanent magnets, because the techniques
developed for synthesizing the metastable phases could overcome the
limitation of the formation of the equilibrium intermetallics. Since the
metastable phases could stabilize under conditions corresponding to a local
minimum of the free energy, the synthesis of the metastable phases provides
us with more possibility for the discovery of new permanent magnets with high
performance. On the other hand, it is interesting to study the formation of the
metastable phases and the transformation between them and the various
equilibrium phases, which is helpful for better understanding the relation of
different crystal structures and the thermodynamic stability of the intermetallics.
In this section, the formation of metastable phases, the transformation
between them and the equilibrium phases, and the magnetic properties of
these phases are investigated.

2.4.1 RFes
The metastable SmFes phase with a Curie temperature of 192C was obtained
by rapid quenching followed by heat-treatment (Miyazaki et aI., 1986, 1988;
Yang et aI., 1986). The Nd-Fe phase with CaCus-type structure was
stabilized by rapid solidification (Stadelmaier et aI., 1986). Structural
transformations from amorphous via metastable to equilibrium phases were
studied for rapidly quenched Sm-Fe-Ti and Nd-Fe-Ti alloys (Xiao et al. , 1995,
1997; Yu et al. , 2000). A complete procession of phase transitions during an
annealing treatment was observed in the rapidly quenched R10 Fego- x Ti x (R =
Sm, Nd) ribbons. X-ray diffraction patterns of Nd,o Fego ribbons at different
annealing temperature (To) are shown in Fig. 2. 6. It is evident that the
CaCus-type structure exists at To = 600 C , the CaCus-type and the TbCur
type structures coexist at To = 700C, the single phase with the TbCurtype
structure exists at To = 750 C, and finally the Th z Znwtype exists at To =
1000C (Yu et al., 2000). The temperature for the emergence (and/or
existence) of the metastable TbCurtype structure in Nd 10 Fego is lower than
that in SmlO Fego (Xiao et al. , 1995. 1997; Yu et al. , 2000). For Nd lO Fego - x Ti x (Yu
et al. 2000). a CaCus-type metastable phase was observed for O~x~lO as
the crystall ization of the amorphous alloys occurs below 700 C. The CaCus-
type structure transforms into a TbCu7-type one with increasing the annealing
temperature (above 700 C ). These melt-spun alloys have different structures
New Rare-Earth Transition-Metal Intermetallic Compounds and ... 65

including Th2Zn,rtype for a< x < 4, TbCu7 and ThMn12 types for 6< x < 1a,
when annealed at high temperature (above 1000 C). The sequence of the
transformation reveals that the CaCus-type phase is a metastable phase with a
high free energy while the TbCu7-type phase is also metastable, but with a
comparatively low free energy.

30 40 50 60 70 80
28(")

Figure 2.6 X-ray diffraction patterns of Nd lO Fe90 ribbons at different annealing


temperature (T.). The arrows denote the a-Fe phase. The triangles denote the
superstructure peaks of the phases studied (After Yu et al. 2000).

2.4.2 RT7 (T = Fe, Co, Ti, etc. )


Liu et al. (1994a, 1994b, 1994c) investigated the formation of the metastable
TbCurtype SmFe7 phase prepared by mechanical alloying and the magnetic
properties of the corresponding nitride SmFe7 Nx The SmFe7 phase formed at
750C transforms gradually into the Th2 Zn,rtype Sm2 Fel7 phase with the
increase of annealing temperature and/or the prolongation of annealing time.
The transformation of SmFe7 into Sm2 Fel7 completes when annealed at
1000 C. The Curie temperature of the SmFe7 phase is always higher than that
of the as-cast main-phase Sm2 Fel7 ( T c = 120C) (L iu et al., 1994a).
Changing of the Curie temperature with annealing temperature confirms that
the structure of the SmFe7 phase is metastable. Mao et al. reported the
existence of the TbCu7 -type RFe7 phase in the whole rare-earth-Fe system
(1996). The lattice parameters a, c and the volume v of the unit cell of
66 Z. D. Zhang

RFe7 -type compounds and their nitrides were given.


The effect of Co substitution for Fe was also studied (Ding et aI., 1992,
1994; Liu et aI., 1994c). Ding et al. found the existence of 1 : 5 and 1 : 7
phases in mechanically alloyed Sm13(Fel-xCox)S7 alloys (Ding et aI., 1992,
1994). The main phase in Sm x C0 1- x powders crystallized at temperatures
between 600'C and 700'C is of TbCu7 -type, and the ordered Th2Zn17 -type
phase forms only after heat treatment at temperatures greater than 700 'C
(Ding et aI., 1993a, 1993b; Liu et aI., 1994c). Sm X Co 1- X alloys made by
mechanical alloying and annealing at T. ~ 700'C exhibit remanence
enhancement of about O. 7 Ms for x = O. 13 and O. 15. The remanence
enhancement, which is ascribed to exchange interactions between
magnetically coupled nanocrystalline phases, increases the maximum energy
product of the alloys. Energy products of 16 - 19 MGOe with coercive forces
between 6 and 9 kOe were obtained for the SmC07 based magnets (Ding et
al., 1993a). Djega-Mariadassou and Bessais (2000) investigated the
emergence of order in nanocrystalline Sm-Fe alloys prepared by mechanical
alloying. It was found that after a high-energy ball-milling and subsequently
annealing, the disordered SmFeg phase with TbCu7-type hexagonal P6/mmm
structure appears as the precursor of the order R3m Sm2 Fe17' Figure 2. 7
shows Curie temperature and hyperfine field versus dumbbell Fe-Fe distances
in nanocrystalline SmFeg alloys with TbCurtype structure (Djega-Mariadassou
and Bessais, 2000). The Fe-Fe distances of the dumbbell configuration in the
ordered Sm2 Fe17 is short with a large negative exchange interaction and thus
any disorder tends to decrease the magnitude of the interaction and increase
the Curie temperature. These results confirm the existence of the TbCu7-type
metastable phase and the transformation to the equilibrium phase Sm2 Fe17 in
the Sm-Fe alloys prepared by mechanical alloying.

210 282
- - - Curie temperature
280
200 - + - Hyperfine field
278
190
276
,.-,
,.-, 180
I...J 2740
-""
;:f 170 272 "5;'
:l::
160 270

150 268
266
140
'---'---'--'--'--..L--'-_'---'---'--l..--'---L--l 264
0.234 0.236 0.238 0.240 0.242 0.244 0.246
dFc.rc(nm)

Figure 2.7 Curie temperature and hyperfine field versus dumbbell Fe-Fe distances in
nanocrystalline SmFeg alloys with TbCu, -type structure (After Djega-Mariadassou and
Bessais, 2000).
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 67

The metastable TbCu7-type disordered structure in Sm2(Co,Mn)17 alloys


was studied by using the rapid quenching technique (Saito et al. , 1988,
1989). The structure of the ribbons changes from Th2Zn17 to TbCu7 -type
structure with increasing the velocity of the roller. The magnetocrystalline
anisotropy constant of the TbCu7 -type disordered phase, experimentally
determined by the singular-point-detection (SPD) method (Saito et al. , 1988,
1989) is of 105-140 kOe, which is 1.2 -1.4 times as large as that (70 - 100 kOe)
with stable Th2Znu-type ordered phase. This increase of the magnetic
anisotropy with changing the crystal structure is caused by the change of the
Sm3+ ion configuration. The Sm3+ ion configuration in the metastable TbCu7-
type structure is completely different from that of the Th2Zn17 -type structure,
concerning two sublattices of Sm ion and dumbbell 3d atoms and the crystal
environment of the Sm3+ ion.
In SmFell Ti alloy ribbons, the structure changes from tetragonal ThMn12-
type to hexagonal TbCu7 -type by changing the roller velocity (Saito et al. ,
1988). The disordered TbCu7 -type phase with Curie temperature of 242 'c
exists for the moderate velocity of the roller, which is between the velocity of
the roller for existence of the ordered Th2Zn17 or ThMn12-type structure and
amorphous.
The as-spun SmlO Fego ribbons consist of an amorphous and a CaCus -type
phase with Curie temperatures of 72 and 218'C (Xiao et al., 1997). The
Curie temperature of the TbCu7-type phase formed when annealed at 850'C is
167 'c. Increasing the annealing temperature from 850 'c to 1000 'c
continuously decreases the Curie temperature of the TbCu7-type phase to
137'C for the Th2Zn17 -type phase. The Curie temperatures of all phases in the
Nd lO Fego ribbons are slightly lower than those of the corresponding ones in the
SmlO Fego ribbons (25, 162, 124 and 86'C, respectively for amorphous,
CaCus -type, TbCu7 -type, and Th 2Zn1rtype phases) (Yu et al. , 2000). The
Th2 Zn17-type Sm-Fe ribbons are magnetically soft whereas the TbCu7-type
ribbons show moderate coercivities up to 2. 14 kOe (Katter et al. , 1991> .
The disordered SmFe7 structure formed in rapidly solidified Sm-Fe and
Sm-Fe-Ti-C alloys were studied by Shield et al. (1998). The addition of Ti
and C results in an order of magnitude grain refinement compared to the binary
alloys. The Curie temperature of the metastable TbCu7-type structure in the
ribbons of Ndg Fe9l-x B x up to 7 at. % B increases both with increasing boron
content and with annealing of the boron-free or low boron ribbons (Gabay
et ai., 1996a, 1996b). The effect of Co substitution for Fe raises the
temperature of the decomposition and the Curie temperature of the TbCu7 -type
phase (Gabay et aI., 1996a). Optimally annealed melt-spun Ndg(Fe,Co)ssB6
alloys have a better temperature stability of the coercivity, but lower
remanence due to the increased residual amounts of the metastable TbCu7-
type phase.
The Curie temperature of 188 'c for SmFe7 phase prepared by HDDR at
the hydrogenation temperature T H = 650 'c decreases to 129 'c of Sm2 Fe17
68 Z. D. Zhang

phase for T H = 850'C (Zhao et al., 1995). The Curie temperature of the
TbCu7-type phase in the Pr-Fe-Mo powders prepared by HDDR is about 200'C
(Jin et aI., 1998). The dehydrogenation at 750 'C results in the complete
formation of the Pr(Fe,Mo)'2 phase. The Curie temperature of the
mechanically alloyed Nd-Fe-Ti powders gradually changes from 180 'c of the
TbCu7-type structure at annealing temperature T. = 650 to 300 'c of the
ThMn 12 -type compound at T. = 1050'C (Jin et al. , 1996, 1997).

2.4.3 RT7 N&

The nitrogenation behavior and thermostability of mechanically alloyed


Sm12.5 Fe87.5 nitride were studied by adjusting the nitrogenation temperature
and time (Liu et aI., 1994a, 1994b, 1994c). A very short time (less than
5 min) is enough for the introduction of the nitrogen atoms into the lattice at
500 'C, because the surface and/or interface of the mechanically alloyed
samples are very fresh. The dependences of the Curie temperature T c on the
annealing temperature for mechanically alloyed Sm x Fel00- x (x = 12. 5, 13. 5)
and the as-cast (x = 10.5) are represented in Fig. 2. 8. The Curie
temperatures of the nitrides are little affected by annealing temperature,
regardless of processing by mechanical alloying or arc-melting. All the RFer
type compounds can absorb a large amount of nitrogen atoms, resulting in a
remarkable lattice expansion of the structure and the enhancement of the Curie
temperature on average, -- 300'C (Mao et aI., 1996). Figure 2. 9 gives Curie

~
'-'

~ ISO SmxFelQo..x
~
S \40 +, x=12.5
B., t;: x=\3.5
' 100'-----'----'----'-----''------'
u 600 700 SOO 900 1000 1100
Annealing temperature CC )
(a)
~ 560
~
.,
:; 520
e.,
P-
4S0 _ _ .J!l. _ _ ->h..--------
.,
;3 44~LOO--7...J.0-0 --SOLO- -900--\--'OO-0--I---'100
..L

Annealing temperature ('C )


(b)

Figure 2. 8 Dependence of Curie temperature T c on annealing temperature (30 min) of


mechanically alloyed Sm12.5 Fes7.5 and Sm135 Fe865 (a) : after being nitrided at 500'C for 3 h(b).
Points at 1100'C are for as cast ingot. (+: X = 12.5, D.: x = 13,5) (Liu et al. , 1994a),
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 69

600
560 .
0:
: RFe7
their nitrides
520
480 0 0
o 0 0
0
440 0
o 0 0

~
'-'
~
200
160


120
80
40
Ce PrNd Sm Gd TbDyHo Er y
R in RFe7

Figure 2.9 Curie temperature of RF~ -type compounds and their nitrides for various rare-
earths (Mao et aI., 1996)"

temperature of RFe7 -type compounds and their nitrides for various rare-earths
(Mao et aI., 1996). The difference among the Curie temperatures of the
various nitrides is less remarkable than among those of the corresponding RFe7
parent compound. The effect of the volume change on the Curie temperature of
R-Fe-N and R-Fe-Ti-Ni compounds with a TbCu7-type structure was studied
(Hu et aI., 1999). The relationship between dlnTc/dlnV and the Curie
temperature T c in rare-earth iron based nitrides with a TbCu7 -type structure
was well described in terms of the itinerant-electron model formal ism"
Figure 2. 10 shows the annealing temperature dependences of intrinsic
coercivity and remanent magnetization for the mechanically alloyed 8m13 FeB7
60 J5

50
0) G'
0 6
640 \0
0 .~
";;:
.~
'2
15<.> 30 + :if
E
.~
1j
~
~
" 20
+ : 47tM,
5
.5 E
A: iHc ~
\0

o 0
600 700 800 900 1000 \ 100
Annealed temperature ('C )

Figure 2.10 Coercivity. remanent magnetization of mechanically alloyed Sm13Fea7NS annealed


at different temperatures. Points at 1100'C are for as-cast ingot (Liu et al. , 1994a)"
70 Z. D. Zhang

annealed at different temperatures and nitrided at 500 C for 30 min (Liu et


al.. 1994a). The results were compared with those of Sm2 Fe17 Nx prepared by
arc-melting. annealing at 1100 C and nitrogenation. The remanence of the
mechanically alloyed Sm13 FeS7 nitride with TbCu7 -type structure changes
slightly with increasing annealing temperature and the intrinsic coercivity
reaches the maximum at 750 C. It is due to the fact that too low annealing
temperature hinders complete solid-state reaction for the formation of a perfect
metastable SmFe7 phase whose nitride is responsible for magnetic hardening.
Annealing at too high a temperature causes excessive grain growth and
evaporation of Sm. which favors the transformation from SmFe7 to Sm2 Fel7 .
The best magnetic properties achieved are of B r = 8. 0 kG. I He = 44. 0 kOe.
(BH ) max = 14. 3 MGOe for nitrided Sm12.5 Fes7.5 alloy (B r denotes the
remancence. 1 He the Loercivity. BH max the maximium energy product). The
high coercivity originates from the contribution of the metastable SmF~ Nx
rather than that of Sm2 Fel7 Nx . The temperature dependences as well as the
thermostability of the magnetic properties of the Sm-Fe-N magnets were
investigated up to 600 'C (Liu et al.. 1994b). The remanence is nearly
unchanged with increasing the anneal ing temperature below 510 C. The
coercivity at 200C of the Sm-Fe-N magnets is still greater than 10 kOe. At
550 C. the main phase SmFe7 Nx decomposes into SmN and ex-Fe. and
consequently the magnetic properties are rapidly degraded.
The hexagonal TbCu7 -type structure forms for NdsFe92- x Ti x alloys
prepared by mechanical alloying and annealing at 650 - 850C (Jin et al.
1996. 1997). Increasing annealing temperature promotes the transformation of
the Nd(Fe. Ti)7 phase to the Nd(Fe. Ti)12 phase. which depends critically on
the annealing temperature and the composition. After nitrogenation. the Curie
temperature of the nano-crystalline Nd-Fe-Ti compounds made by mechanical
alloying significantly increases to above 430 C. The structure and phase
formation of mechanically alloyed Rll FeS4 Ti 5 with R = Nd and Sm and their
nitrides were also studied by Ding et al. (1995).
Sm-Fe-N magnets were prepared by rapid solidification (Christodoulou
and Takeshita. 1993a; Katter et al . 1991). The hexagonal TbCu7 -type phase
shows up in melt spun ribbons of Sm-Fe alloys with stoichiometry SmFeg. only
at wheel velocities above 15 m/s (Katter et al.. 1991). At higher Sm
concentrations or lower quenching rates the structure changes to the Th2Zn 17-
type. Sm-rich alloys solidified at high quenching rates (55 m/s) consist of a
mixture of metastable TbCu7-type SmFe7 and an amorphous matrix
(Christodoulou and Takeshita. 1993a. 1993b). Subsequent heat treatment at
650C for 10 min form Th 2Zn17-type Sm2 Fell phase in metastable equilibrium
with MgCU2 -type SmFe2 phase.
The Curie temperatures of the two metastable phases in the rapidly
quenched SmlO Fego- x Ti x alloys are increased after nitrogenation. The Curie
New Rare-Earth Transition-Metal Intermetallic Compounds and ... 71

temperature for TbCu7 -type Sm(Fe, TD 7Nx is 470 "c without significant
dependence on Ti content, compared with 431C for CaCus -type
Sm(Fe, TDsN x (Xiao et a!., 1997). The best hard magnetic properties were
obtained after nitrogenation for x = 3 in the rapidly quenched SmlO Fe90- x Ti x
ribbons (Xiao et a!., 1997). For the CaCus-type one, I H c = 8.6 kOe, B, =
8. 4 kG and (BH) max = 10. 2 MGOe and for the TbCu7 -type phase, I He =
17.6 kOe, B,= 7.7 kG and (BH)max = 6.8 MGOe. However, for Nd-Fe-Ti
alloys, the magnetic properties of the metastable Nd (Fe, Tj)s Nx and Nd (Fe,
TD 7Nx phases are magnetically soft (Yu et a!. , 2000).
Nitrogenation increases the Curie temperatures to 470C and the
saturation polarizations to 1.4 and 1.51 T for the TbCur and the Th2Zn17 -type
phases in the Sm-Fe ribbons, respectively (KaUer et a!., 1991). The best
hard magnetic properties for isotropic TbCu7-type Sm-Fe-N magnets were
obtained for quenched, annealed, nitrided SmlO.6 Fe89.4 with a coercivity of
6.2 kOe, a remanence of 8.6 kG and an energy product of 8.7 MGOe. The
best result of the coercivity I H c = 23.7 kOe was achieved for Sm12.S Fe87.5 Nx
quenched at 50 mis, annealed at 700 "c for 30 min, and then nitrogenated at
460"c for 4 h (Kim et a!., 1997).
Structured and magnetic properties of Sm-Fe-N magnets fabricated by
hydrogenation and nitrogenation processes were investigated (Christodoulou
and Kateshita, 1993b; Sugimoto et a!., 1992; Zhao et a!. , 1995). The role
of hydrogen in the Sm2 Fe17-based alloys is similar to that in the
Nd2Fe14 B-based alloys. The hydrogenation at 600C leads to the
disproportionation reaction of hydrogen with Sm2 Fe17 , resulting in the
formation of SmH2+ x (x < 0 and a-Fe phases. After the disproportionation,
the desorption and recombination occur almost simultaneously at about 650 "c
under vacuum or argon atmosphere. During the desorption-recombination,
SmH2+ x dissociates into elemental Sm and hydrogen gas and then the Sm
recombines with ex-Fe to form the metastable SmFe7 phase with a TbCu7-type
structure. With increasing temperature, the desorption-recombination tends to
completion and the metastable SmFe7 phase transforms gradually into the
Sm2 Fe17 phase that begins to occur when hydrogenated at 800 C. The Curie
temperatures of SmFe7 Nx and Sm2 Fel7 Nx phases prepared by the processes of
above HDDR and nitrogenation are all approximately 480C (Zhao et al. ,
1995). The completion of the desorption-recombination reaction is beneficial
to the magnetic properties of the isotropic Sm-Fe-N magnets. The best results
achieved at the optimum hydrogenation temperature of 800 C are: B, =
8.1 kG, ,H c = 21.0 kOe, (BH)max= 13.0 MGOe. A too high hydrogenation
temperature degrades the magnetic properties, due to the growth of the grains
and the vaporization of samarium. The effect of Ti or Zr substitution for Fe on
the structure and magnetic properties of the Sm-Fe-N magnets prepared by
HDDR was studied (Zhao et a!., 1999; Gebel et ai., 1997) (see Fig. 2.10.
72 Z. D. Zhang

HDDR-treated and nitrogenated Sm,O.SFe88.SZrl.ONy (y -.. 16) has the


properties of ,He = 31 kOe and (BH)max = 12.9 MGOe (Gebel et al. 1997).

20
~

0'"
o 15
6
g
!10
~

0'"
C 5
:zt
0
0 0.Q2 0.04 0.06 0.08 0.10 0.12
Ti content, x

Figure 2. 11 Composition dependence of magnetic properties of the Sm-Fe-Ti-N magnets


prepared by HDDR process (Zhao et al.. 1999).

2. 4. 4 R-Fe-C
The changes of the magnetic properties of the R-Fe-C alloys with annealing
temperature and composition correlate directly with the formation of the hard
magnetic RzFe14C phase (Sui et al.. 1997a. 1998a). A status diagram of the
transformation of main phases is illustrated in Fig. 2.12. depending on annealing
temperature and Nd content for mechanically alloyed NdyFel00-1.Sy Co. Sy alloys
annealed for 35 min. The temperature range for the formation of NdzFe14 C

1000 L

~ 900
~
~ 3-
e 3- ~ 3- eo+
:J
~ + +
eo cI>
e'" 800
0 ..
B H

E
'"
f-o
cI>+L
700

H'I'
600
11 12 13 14 15 16 17 18
Nd content (at.%)

Figure 2. 12 Status diagram of the transformation of main phases depending on annealing


temperature and Nd content for mechanically alloyed NdyFelOo-15yC05y alloys annealed for
35 min. <D = Nd2 Fe" C, L = N~ Fell C x ,'Y = ex-Fe (Sui et al. , 1998a).
New Rare-Earth Transition-Metal Intermetallic Compounds and .. 73

phase becomes wider in the case of mechanical alloying, because of the


metastable status of the system. The coercivity of i He = 7. 5 kOe was
obtained in mechanically alloyed Nd 16 Fe7S.a Ca.z annealed at 900 C, which
contains a certain amount of NdzFe17 Cx ' The best magnetic properties
achieved in Nd16Fe76BsC3 alloy are: ,He = 18.6 kOe, B, = 7.1 kG and
(BH)max = 11.5 MGOe. An fcc structure rare-earth-rich phase, which might be
non-magnetic or paramagnetic, exists in the mechanically alloyed R-Fe-C
alloys, which is beneficial to high coercivity because the homogeneous
distribution of the phase separates the hard magnetic grains. The highest
coercivities obtained in the Dy-Fe-C system are 93 kOe (Sui et aI., 1997c, 1997d).
A process consisting of mechanical alloying and re-milling was developed
to prepare the Sm-Fe-C based magnets (Geng et al. , 1999, 2000). In order
to achieve better magnetic properties, it is necessary to avoid the formation of
the Sm Z Fe 14 C phase and to increase the carbon content in the Sm Z Fe 17 C x
phase. The carbon atoms in the alloys diffuse very slowly at T. < 650 C .
Increasing the annealing temperature is not helpful for increasing the carbon
content. The procedure of re-milling for a short time and re-annealing at lower
temperatures overcomes the difficulties above for the formation of Smz Fe17Cx
with high C content (Geng et aI., 1999, 2000). Upon the re-milling, more
interfaces and/or the grain boundaries appear while the carbon atoms are
mixed efficiently with the grains of the SmZ Fe 17 Cx phase. During re-annealing
at low temperatures, the carbon atoms can diffuse easily into the lattice of the
Sm Z Fe 17 Cx phase. The extra energy provided by the re-milling is responsible
for overcoming the energy barriers of the diffusion. The grain size after re-
annealing is rruch smaller than that before the re-milling procedure. Figure 2. 13
shows hysteresis loops of MA Sm-Fe-C alloys annealed at 850 "C for 30 min,
and subsequently remilled for 1 hand re-annealed at 600 C for 30 min. The

10
8

(2)

6 8 10

::::::::::~::==::::"':-8
-10
Applied field (kOc)

Figure 2. 13 Hysteresis loops of MA Sm-Fe-C alloys (1) annealed at 850'C for 30 min and
(2) subsequently remilled for 1 hand re-annealed at 600'C for 30 min (Geng et al. , 2000).
74 Z. D. Zhang

process of remilling and re-annealing can evidently improve the magnetic


properties of the alloys. The best magnetic properties of the Sm-Fe-C alloys
prepared by the procedure of re-milling and re-annealing are of Sm14 Fe79 C7
reannealed at 600C for 35 min, B s = 11.8 kG, B, = 9.53 kG, tHe = 2.67 kOe,
(BH)max = 10.7 MGOe (Geng et al. , 1999).
The composition dependence of the magnetic properties of the annealed
Nd-Fe-C flakes is consistent with the phase analysis (Coehoorn et al. , 1989).
High coercivities were observed along the central axis of the region in which
Nd2Fe14 C formed after annealing. Because the region for the formation of
Nd2Fe14 C is quite narrow, there is also only a narrow region for the high
coercivity field. The highest coercivity was 11.6 kOe for Nd14.sFe76.4Cs.s,
while very low coercivity (less than 1 kOe) was obtained for the sample
without containing Nd2Fe 14 C.
The intrinsic coercivity of the Nd 16 Fe75 C9 alloy processed by HDDR is
8.0 kOe, which is higher than that of the master alloy made by mechanical
alloying (Sui et aI., 1998b). This is because of the existence of more Nd-rich
phases, leading to more complete separation of the Nd2Fe'4 C grains that
increases the expansion field of the reversed domain and decreases the
magnetostatic- coupling field between the hard magnetic grains.

As shown in the last section, the chemical bond of boron nitride can be broken
down by arc-melting (Zhang et al. , 1999) to allow the combination of Band N
with rare-earth and transition-metal atoms to form the Nd2Fe14 B-type structure.
By using mechanical alloying, we successfully broke the bond of boron nitride
BN to combine with Nd and Fe to form Nd2Fe14 BN s alloys with I He = 20.0 kOe
(Liu et al., 1999, 2000a). The decomposition of pyrolytic boron nitride (p-
BN) during milling was first studied as a function of the milling time. It was
found that the p-BN compound could be partially decomposed by milling until an
amorphous p-BN phase is formed so that the content of nitrogen in the p-BN
system will not continue to change upon the milling. A Nd2Fe14 BN s phase with
x up to 0.25 coexists with some amounts of NdN, Nd-rich phase and a-Fe.
Curie temperature of the Nd2Fe14BNo.25 phase is 335 C, slightly higher than
that of the Nd2Fe14B compound. A pre-milling process of p-BN favors the
formation of the Nd2Fe14 BN s phase. The magnetic properties of Nd 16 Fe76 BsNs
alloys prepared by using pre-milled p-BN are better than those made of non-
pre-milled p-BN. Figure 2. 14 shows hysteresis loops at room temperature of
four Nd2Fe14 BN s alloys annealed at 750C for 30 min. A coercivity higher than
20 kOe was achieved for Nd2 Fe14 BN s based alloys by adding excess Nd,
which is close to that of Nd 16 Fe76 Bs prepared by using pure B. Effects of Nd
and N contents on structure and magnetic properties were investigated for Nd-
Fe-B-N alloys prepared by mechanical alloying (Liu et al. , 2000a, 2000b).
The nitrogen contents in the interstitial compounds Nd2Fe14 BN s and also in the
New Rare-Earth Transition-Metal Intermetallic Compounds and . 75

alloys can be controlled to some extent by changing the Nd content (Liu et al.
2000a). Curie temperature of mechanically alloyed NdxFe92-xBsNy annealed
at 750'C for 30 min decreases from 338 "C for x = 16 to 295'C for x = 30
with increasing Nd contents. The decrease of the Curie temperature is due to
the decrease of the content of nitrogen atoms that enter the lattice of
Nd2Fel4B-type structure. accompanying the formation of a Nd-rich phase (Liu
et al . 2000b).

10

s
g
c
o
.~ 0
.~

6h
~ -s

-40 -20 0 20 40
Applied field (kOe)

Figure 2. 14 Hysteresis loops at room temperature of the Nd-Fe-B-N samples MA, MA2
MA3 and MA4 annealed at 750'C for 30 min (Liu et al.. 1999). The sample preparation
and the phase identification for the MAl' MA2 MA3 and MA4 samples were summarized in
Table 1 of (Liu et al.. 1999).

2.4.6 RFe12
The ThMnl2-type RFe12 phase does exist in its non-equilibrium state. without
stabilization of third elements. such as Ti. V. Mo. Cr. Mn. etc. The
metastable SmFe12 films with a ThMn12 -type tetragonal structure were
successfully deposited by sputtering (Cadieu et al.. 1991; Wang et al.
1993). Figure 2. 15 shows a hysterisis loop for SmFe12 Nx film measured at 5 K
with the direction of a magnetic field either parallel or perpendicular to the film
(Wang et al.. 1993). The binary SmFe12 and Sm(Fe. n 12 films with T = Ti
and V crystallize in the ThMnl2-type structure with the c axes perpendicular to
the film plane (Cadieu et al.. 1991). The Curie temperature for the SmFe12
film is about 320 'C which is similar to that for a bulk SmFell Ti sample (Wang
et al. 1993). The anisotropy field was estimated to be about 85 kOe for the
SmFe12 film after subtracting the demagnetization field by using a de-
magnetization factor 41T and the bulk value of the saturation magnetization
(Wang et al.. 1993). The anisotropy field decreases substantially upon
nitrogenation. in accordance with the transition of the anisotropy from uniaxial
76 Z. D. Zhang

to planar observed in the SmFe based pseudo-binary bulk compounds with the
ThMn 12 -type structure. The measured in-plane coercivity for the SmFe12 film at
room temperature is about 3.5 kOe (Wang et aI., 1993). The coercivity for
nitrided film is lower than that of the SmFe12 film but still high considering a
strictly planar anisotropy system. This is likely due to some residual TbCu7
and/ or Th 2Zn17 phases that could be strongly exchanged-coupled with the
majority ThMn12 -type phase and lead to an overall increase of the coercivity.
However, Sm(Fe,T)12 films with T =
Ti and V were synthesized successfully,
which exhibit a very dominant (002) texture with intrinsic coercivity of 5 or
more kOe, flux density B s of more than 10 kG perpendicular to the film plane
(Cadieu et aI., 1991). No discernible columnar structure is exhibited so that
the static energy product measured perpendicular to the film plane is nearly
21 MGOe (Cadieu et aI., 1991). It is still possible to form a well-defined
SmFe12 phase as the third-element addition is reduced to zero, which can
exhibit extreme perpendicular anisotropy magnetic properties.

1.0

0.5
t:
o
.~
.~

~ 0 T=S K
E
]
"T;l
~ --{l.S

-1.0
'--_-'--_---'---_----"--_-----'-_----l_--'
-60 -40 -20 0 20 40 60
Field (kOe)

Figure 2. 15 Hysterisis loop for SmFe'2 Nx film measured at 5 K with the direction of a
magnetic field either parallel or perpendicular to the film (Wang et al. 1993).

2. 5 Discussion

The manufacture of the super-magnets with better magnetic properties than the
Nd2Fe14 B-based magnets would be a challenge to the scientists and engineers
in the field of the permanent magnets in the new century. In order to discover the
new permanent magnetic materials with high performance, unceasing efforts
have been made mainly along the following three directions: CD search new rare-
earth transition metal intermetallics; (2) synthesize the rare-earth metastable
phases; @design and prepare the nanocomposite magnets.
New Rare-Earth Transition-Metal Intermetallic Compounds and ... 77

Researching new rare-earth transition metal intermetallics has met some


difficulties: First, the formation of the stabilized R-T intermetallics is limited
by atomic size, binding energy, the crystallographic symmetry, etc. These
factors limit the numbers for the type and the quantity of the rare-earth
transition metal intermetallics. Almost all of the binary rare earth transition
metal systems/ compounds have been uncovered. Although the ternary and the
multi-component compounds have been relatively unexplored, it is believed
that there is a long way to go before the breakthrough of finding a novel multi-
component compound with high performance. Second, the contributions to the
intrinsic magnetic properties of different elements in one compound are usually
constrained to each other. It is well-known that a compound which is suitable
for application of permanent magnets must possess high saturation
magnetization, high easy axis magnetocrystalline anisotropy and high Curie
temperature. Because the property of the elements that are necessary for
stabilizing the compound is different, the contributions of the elements may
cancel out each other, and because of the interaction of electrons, the good
property of one element may be suppressed by the introduction of other elements.
Besides the preparation of the nanocomposite magnets, the synthesis of
the rare-earth metastable phases is also one of the main directions in the area
of the rare earth permanents. The metastable phase is not thermodynamically
the most stable phase, but rather the temporarily stable one in a certain
condition. The formation of the metastable phase corresponds to the local
minimum of the free energy, which can be produced at various extreme non-
equilibrium conditions. As mentioned above, mechanical alloying, rapid
quenching, HDDR, solid-state reaction, solid-gas reaction and sputtering are
among the technologies for preparation of the rare-earth metastable phases.
Therefore, the formation of the metastable phases could break the Iimitation of
the phase equilibrium conditions of the alloy systems and of the
crystallographic symmetry. The metastable phases can have different
structures and compositions with those of the equilibrium phases, and may
have good magnetic properties. Furthermore, the study of the metastable
phases tries to achieve an understanding of the mechanism of the
transformation between them and the equilibrium phases, which is beneficial to
achieve an understanding of the crystallographic characters and the stability of
the rare earth transition metal intermetall ics.
If more than one metastable phase exists in certain conditions, the energy
barrier between the local energy minima (and also the lowest energy
minimum) hinders the formation of other metastable phases (and also the
equilibrium phase) dynamically. If the thermal energy would overcome the
energy barrier, the new metastable phase would form when annealing the as-
prepared metastable materials at high temperatures. However, the
temperatures for the transformation of metastable phases differ for different
synthesis methods. The thermostability of the metastable phase depends on
the preparation process and the composition of the alloys.
It is interesting to discuss, from the point view of symmetry, the
78 Z. D. Zhang

formation, stabilization and transformation of the rare-earth metastable


phases. The symmetry of the amorphous phase is highest, which retains the
disordered structure of the liquid phase at high temperatures. The symmetry of
the phase with the hexagonal structure is higher than that with the tetragonal
and the rhombohedral ones. The hexagonal CaCus-type structure has the
highest symmetry among those of the rare-earth metastable phases and the
equilibrium ones because it is the basic one from which the structures for
various R-T compounds can be derived. Comparing two phases with hexagonal
structures derived from the CaCus-type, the one with disordered replacement
has higher symmetry. Thus the symmetry of the TbCu7 -type structure is the
highest among all the structures with the exception of the CaCus-type. The
symmetry of the tetragonal structure is higher than that of the rhombohedral
one. The symmetry of a monoclinic structure is lowest among the structures
above. We found the sequence of their symmetries for the structures of the R-
T compounds as follows (Zhang et al. , 2000): Amorphous, CaCus-type with
space group P6/mmm, TbCu7-type with space group P6/mmm, Th2Ni 17 -type
with space group P6 3 /mmc, Nd2Fe14 B-type with space group P4 2/mnm,
ThMn12-type with space group 14/mmm, Th2Zn17-type with space group R3m
and Nd3 (Fe, Tj)29-type structure with space group Adm.
The differences between the symmetries of the phases lead to the
differences on the local minima of the free energy for the formation of the
phases. However, many other factors, such as the composition, the binding
force and/or the binding energy when the atoms combine, also play an
important role for the formation and the stability of the phases. Usually, the
difference between the energy minima for the formation of the last five
equilibrium phases in the subsequence may be quite small, depending
sensitively on the composition of alloys and the condition of the synthetic
processes. Therefore, it is always true that the complete sequence of the
phases listed above may not occur during the phase transformation. In
general, only one equilibrium phase forms as the ending phase for a certain
condition of compositions and processes. It has been found that the substitution
Ti (or V, Mo, W, Cr, Mn, Si, etc.) for Fe favors the formation of the
ThMn2 -type structure, whereas addition of B (or C) prefers the formation of
the Nd2 Fe14 B-type structure (or the Th z Zn17 -type one). Co addition may
stabilize the formation of the CaCus-type structure and/or the TbCu7-type one.
This results in the existence of the different phase transformations for various
R-T systems as well as various processes.
Because the CaCus -type structure with a high free-energy and a high
crystallographic symmetry is the basic one of the R-T phases, it is expected
that it may exist in all the R-T systems in either metastable or equilibrium
state. At the temperatures close to the crystallization temperature of the
amorphous phase, the CaCus -type structure may start to form in the
metastable state. However, for most of the R-T systems, the temperature
interval for the formation of the metastable CaCus -type phase may be very
narrow so that it is hard to be observed experimentally. The existence of the
New Rare-Earth Transition-Metal Intermetallic Compounds and. . . 79

CaCus-type metastable phase can be observed by a careful experiment with a


very slow heating rate and a short annealing time. Fortunately, the
temperature interval for the existence of the TbCu7 -type metastable phases is
wide enough for observing it in more R-T systems prepared by various
processes. The order-disorder transformation between the TbCu7-type
metastable phases and others would occur if both of the following conditions
would be satisfied: CD the relative atomic positions in the structure of the
disordered phase should remain undisplaced in the ordered structure (Lipson,
1950); (g)the symmetry of the order structure should be lower than, or at least
equal to, that of the disordered structure (Lipson, 1950; Stoloff and Davis,
1966; Barrett and Massalski, 1966; Pearson, 1972).
It is expected that many more efforts will put on synthesizing more multi-
component rare-earth intermetall ic compounds and/or metastable phases,
which benefit from the rapid advances in the synthesizing processes. The
powerful methods, such as the mechanochemical process, the mechanical
grinding, sputtering, laser processing, ion-beam bombardment, and various
epitaxial processes, will become more popular for searching the novel hard
magnetic metastable phases. Furthermore, since the metastable phases
usually crystall ite in the grain size of nanometers, the demarcation line
between the research directions of the metastable phases and the
nanocomposite permanent magnets will become dim. The interpenetration of
the three main directions is a tendency in the area of the rare-earth permanent
magnets.

2.6 Conclusions

The recent advances in intermetallics and metastable phases in rare-earth


permanent magnets have been reviewed. One of the main trends in this field is
to stabilize the new intermetallics with multi-component compositions. The
best possibilities for searching compounds for good magnets are in the multi-
component systems with magnetic rare-earths, magnetic transition-metals and
other kinds of elements, like small atoms. The crystal structure of the
compounds should be either tetragonal, hexagonal or rhombohedral for a large
magnetocrystalline anisotropy. If one succeeded in changing the law of the
exchange interactions between some magnetic ions, for example, aligning
parallel the magnetic moments of the heavy rare-earths and transition-metals,
the magnetic moment together with huge magnetocrystalline anisotropy of the
heavy rare earths would result in a novel compound for a supermagnet. This
might be realized in some limit cases, such as synthesizing in high press or
combining with some special elements. Adjusting the lattice parameters
through adding new atoms might alter the magnetic behaviors dramatically so
that the antiferromagnetic coupling of some transition metal ions might become
80 Z. D. Zhang

the ferromagnetic one. while the easy plane magnetic anisotropy might
become the easy axis one. Another possibility is to enhance the magnetic
moments of the transition-metals by changing the band structure of the
compounds. It is believed that the space for searching the new compounds
with good magnetic properties does exist.
On the other hand. various metastable phases possess good intrinsic
magnetic properties showing great potential in applications of permanent
magnets. Various processes. including mechanical alloying. mechanical
milling. rapid quenching. HDDR. solid-state reaction. solid-gas reaction. and
sputtering can synthesize these metastable phases. The relations between the
structures of the metastable phases and the equilibrium ones and the
transformations from the metastable phases to the equilibrium ones have been
studied systematically. The difference between the symmetries of the
metastable phases and the equilibrium ones is found to be one of the main
factors for the formation and the transformation of the phases. besides the
binding energy. Of course. the procedure of the synthesis and the composition
of the alloys may also control these processes. The study of the metastable
phases opens a new way for searching and synthesizing rare-earth permanent-
magnet materials.

References
Allemand.J . A. Letant. J. M. Moreau. J. P. Nozieres and R. Perrier de la
Bathie. J. Less-Common Met. 166: 73( 1990)
Barrett. C. S. and T. B. Massalski. Structure of Metals. 3rd ed. McGraw-
Hill. New York.p.270(1966)
Benjamin. J.S. Metall. Trans. 1: 2943(1970)
Benjamin. J.S. Sci. Am. 234: 40(1976)
Bouchet. G. J. Laforest. R. Lemaire and J. Schweizer. C. R. Acad. Sci.
Par~. 262: 1(1966)
Bulcock. S. Heng Zhang. E. Wu and S. J. Campbell. J. Mater. Sci. Lett.
17: 1791(1998)
Burzo. E. Rep. Prog. Phys. 61: 1099( 1998)
Buschow. K. H. J. Phys. Status Solidi. A 7: 199 ( 1971 )
Buschow. K.H.J. Rep. Prog. Phys. 40: 1179(1977)
Buschow. K. H. J. In: E. P. Wohlfarth ed. Handbook on Ferromagnetic
Materials. Vol. 1. North-Holland. Amsterdam. p.297(1980)
Buschow. K. H. J. In: E. P. Wohlfarth and K. H. J. Buschow. ed. Handbook
on Ferromagnetic Materials. Vol. 4. North-Holland. Amsterdam. p. 1( 1988)
Buschow. K. H. J. D. B. de Mooij and C. J. M. Denissen. J. Less-Common.
Met. 141: L15(1988)
Buschow. K. H. J. Rep. Prog. Phys. S4: 1123C 1991)
Buschow. K. H. J. In: Buschow K. H. J. ed. Handbook on Magnetic Materials.
Vol. 10. Elsevier Science. B.V . Amsterdam. p.463C1997>
Cadieu, F. J. , H. Hegde, A. Navarathna. R. Rani and K. Chen. Appl. Phys.
New Rare-Earth Transition-Metal Intermetallic Compounds and ... 81

Lett. 59: 875(1991)


Campbell, S.J., Heng Zhang, H.S. Li and E. Wu. J. Magn. Magn. Mater.,
177-181: 1103( 1998)
Chen, K., H. Hegde, S.U. JenandF.J. Cadieu. J. Appl. Phys. 73: 5923(1993)
Christodoulou, C.N. andT. Takeshita. J. Alloys and Compounds 196: 161C1993a)
Christodoulou, C. N. and T. Kakeshita. J. Alloys and Compounds 196: 155(1993b)
Chuang, Y. C. , Dan Zhang, Tong Zhao, Zhi-dong Zhang, X. K. Sun and F. R.
de Boer. J. Alloys and Compounds 221: 60 ( 1995)
Coehoorn, R. , J. P. W. B. Duchateau and C. J. M. Denissen. J. Appl. Phys.
65: 704(1989)
Coey, J.M.D and Hong Sun. J. Magn. Magn. Mater. 87: L251(1990)
Coey, J. M. D, Hong Sun, Y. Otani and D. P. F. Hurley. J. Magn. Magn.
Mater. 98: 76(1991)
Coey, J. M. D., O. N. Oi, K. G. Knoch, A. Leithe-Jasper, P. Rogl. J.
Magn. Magn. Mater. 129: 87(1994)
Coey, J.M.D. J. Magn. Magn. Mater. 140-144: 1041(1995)
de Boer, F. R., Ying-kai Huang, D. B. de Mooij and K. H. J. Buschow. J.
Less-Common Met. 135: 199 ( 1987)
de Boer, F.R., Ying-kai Huang, Zhi-dong Zhang, D.B. de Mooij and K.H.J.
Buschow. J. Magn. Magn. Mater. 72: 167(1988)
de Groot, C. H. , K. H. J. Buschow and F. R. de Boer. Phys. Rev. B 57:
11472(1998)
Ding, J. , P. G. McCormick and R. Street. J. Alloys and Compounds 189: 83
(1992)
Ding, J. , Y. Liu, P. G. McCormick and R. Street. J. Magn. Magn. Mater.
123: L239(1993a)
Ding, J., P. G. McCormick and R. Street. J. Alloys and Compounds 191: 197
(1993b)
Ding, J., P. G. McCormick and R. Street. J. Magn. Magn. Mater. 135: 200
(1994)
Ding, J., P. G. McCormick and R. Street. J. Alloys and Compounds 217: 108
(1995)
Djega-Mariadassou, C. and L. Bessais. J. 'Magn. Magn. Mater. 210: 81
(2000)
Dong, X. L. , Z. D. Zhang, H. T. Kim, D. Zhang, X. C. Kou, W. Liu, X. K.
Sun, Y.C. Chuang, B.S. Zhang, and H.N. Du. J. Phys. Condo Matt. 8:
6079(1996)
Eisses, J., D. B. de Mooij, K. H. J. Buschow and G. Martinek. J. Less-
Common. Met. 171: 17 ( 1991)
Florio, J. V. , R. E. Rundle and A. I. Snow. Acta Crystallogr. 5: 449 ( 1952)
Franse, J. J. M. and R. J. Radwanski. In: K. H. J. Buschow, ed. Handbook of
Magnetic Materials, Vol. 7. Elsevier, Amsterdam, p. 307 (1993)
Fujii, H. and H. Sun. In: Buschow K. H. J., ed. Handbook on Magnetic
Materials, Vol. 9. Elsevier Science B.V., Amsterdam, p.304(1995)
Gabay, A. M., A. G. Popov, V. S. Gavico, Yeo V. Belozerov, A. S.
82 Z. D. Zhang

Yermolenko and N. N. Shchegoleva. J. Alloys and Compounds 237: 101


( 1996a)
Gabay, A. M., A. G. Popov, V. S. Gavico, Yeo V. Belozerov and A. S.
Yermolenko. J. Alloys and Compounds 245: 119(1996b)
Gebel, B., M. Kubis and K. H. MOiler. J. Magn. Magn. Mater. 174: L1
( 1997)
Geng, Dian-yu, Zhi-dong Zhang, Bao-zhi Cui, Zhi-jun Guo, Wei Liu, Xin-guo
Zhao, Tong Zhao and Ji-wen Liu. J. Alloys and Compounds 291: 276(1999)
Geng, Dian-yu, Zhi-dong Zhang, Bao-zhi Cui, Zhi-jun Guo, Wei Liu and Xin-
guo Zhao. J. Appl. Phys. 87: 5296(2000)
Hegde, H. X.R. Qian, J.G. AhnandF.J. Cadieu. J. Appl. Phys. 79: 5961
(1996)
Herbst, J.F. Rev. Mod. Phys. 63: 819(1991)
Hu, B. P. , J. M. D. Coey, H. Klesnar and P. Rogl. J. Magn. Magn. Mater.
117: 225(1992)
Hu, B.P., G.C. Liu, Y.Z. Wang, B. Nasunjilegal. N. Tang, F.M. Yang,
H. S. Li and J. M. Cadogen. J. Phys. Condensed Matter. 6: L595(1994)
Hu, J.F., K.Y. Wang, Y.Z. Wang, B.P. HuandZ.X. Wang. J. Alloys and
Compounds 292: 233( 1999)
Jin, z.a., x. K. Sun. W. Liu, X. G. Zhao, Q. F. Xiao, Y. C. Sui, Z. G.
Wang. Z.D. Zhang and H.X. Tan. J. Appl. Phys. 79: 5525(1996)
Jin, Z. Q. , X. K. Sun, W. Liu, X. G. Zhao, Q. F. Xiao, Y. C. Sui, Z. G.
Wang, Z. D. Zhang and H. X. Tan. J. Magn. Magn. Mater. 169: 135
( 1997)
Jin, Z. Q. , W. Tang, J. R. Zhang. L. Y. Lu, S. L. Tang and Y. W. Du. J.
Magn. Magn. Mater. 187: 231(1998)
Kalogirou, O. , V. Psycharis, L. Saettas and D. Niarchos. J. Magn. Magn.
Mater. 146: 335(1995)
Katter. M. , J. Wecker and L. Schultz. J. Appl. Phys. 70: 3188 ( 1991)
Khan Y . Acta Crystallogr. B 29: 2502(1973)
Kim. H. T . Q. F. Xiao, Z. D. Zhang, D. Y. Geng, Y. B. Kim. T. K. Kim and
H.M. Kwon. J. Magn. Magn. Mater. 173: 295(1997)
Kou, X. C. , R. Grossinger, X. Li, J. P. Liu, F. R. de Boer. M. Katter, J.
Wecker. L. Schultz. T. H. Jacobs and K. H. J. Buschow. J. Appl. Phys.
70: 6015(1991)
Li, H.S. andJ.M.D. Coey. In: K.H.J. Buschow,ed. Handbook of Magnetic
Materials. Vol. 6. Elsevier. Amsterdam, p. 1(1991)
Lipson, H. Prog. Metal. Phys. 2: 1(1950)
Liu, N. C . H. H. Stadelmaier and G. Schneider. J. Appl. Phys. 61: 3574
( 1987)
Liu. W. , Z. D. Zhang. X. K. Sun, Y. C. Chuang. F. M. Yang and F. R. de
Boer. Solid State Commun. 76: 1375( 1990)
Liu. J.P., Y. Liu, R. SkomskiandD.J. Sellmyer. J. Appl. Phys. 85: 4812
(1999)
Liu, Wei, Qun Wang, X.K. Sun, Xin-guo Zhao, Tong Zhao, Zhi-dong Zhang
New Rare-Earth Transition-Metal Intermetallic Compounds and . . . 83

and Y.C. Chuang. J. Magn. Magn. Mater. 131: 413(1994a)


Liu, Wei, Qun Wang, X. K. Sun, Xin-guo Zhao, Qun-feng Xiao, Zhi-dong
Zhang, Tong Zhao and Y. C. Chuang. J. Alloys and Compounds 215: 257
(1994b)
Liu, Wei, Qun Wang, X. K. Sun, Xin-guo Zhao, Qun-feng Xiao, Zhi-dong
Zhang and Y. C. Chuang. Solid State Commun. 91: 971( 1994c)
Liu, W., Z. D. Zhang, X. K. Sun, J. F. He and X. G. Zhao. J. Phys. D
Appl. Phys. 32: 1591( 1999)
Liu, W. , Z. D. Zhang, J. P. Liu, X. K. Sun and D. J. Sellmyer. J. Appl.
Phys. 87: 5332(2000a)
Liu, W., Z.D. Zhang, J.P. Liu, X.K. Sun, J.F. He, H. Tang, B.Z.Cuiand
X. G. Zhao. J. Alloys and Compounds 309: 172(2000b)
Mao, O. , Z. Altounian, J. Yang and J. O. Stem-Olsen. J. Appl. Phys. 79:
5536(1996)
McCormick, P. G. , J. Ding, E. H. Feutrill and R. Street. J. Magn. Magn.
Mater. 157-158:,7(1996)
McCormick P. G. In: K. A. Gschneidner, Jr. and L. Eyring, eds. Handbook on
the Physics and Chemistry of Rare-Earths, Vol. 24. Elsevier Science B.
V., Amsterdam, Chapter 160. p.47(1997)
Miyazaki, T., Xing-bo Yang, K. Takakura and M. Takahashi. J. Magn.
Magn. Mater. 60: 211(1986)
Miyazaki, T. , M. Takahashi, Xing-bo Yang, H. Saito and M. Takahashi. J.
Appl. Phys. 64: 5974 ( 1988)
Pan, H., Y. Chen, X. Han, C. Chen and F. Yang. J. Magn. Magn. Mater.
185: 77(1998)
Pearson, W. B. The Crystal Chemistry and Physics of Metals and Alloys.
Wiley, New York, p. 16 and p.643(1972)
Ray, A. E. Acta Crystallogr. 21: 426 ( 1966)
Sagawa, M. , S. Fuj imura, M. Togawa and Y. Matuura. J. Appl. Phys. 55:
2083(1984)
Saito, H. , M. Takahashi and T. Wakiyama. J. Appl. Phys. 64: 5965(1988)
Saito, H., M. Takahashi, T. Wakiyama, G. Kido and H. Nakagawa. J.
Magn. Magn. Mater. 82: 322(1989)
Schultz, L. , J. Wecker and E. Hellstern. J. Appl. Phys. 61: 3583( 1987)
Schultz, L. and J. Wecker. J. Appl. Phys. 64: 5711(1988)
Shcherbakov Y. V. , G. V. Ivanova, A. S. Yermolenko and V. S. Gaviko. J.
Alloys and Compounds 182: 199( 1992)
Shield, J.E., C.P. Li and Branagan.D.J. J. Magn. Magn. Mater. 188: 353
(1998)
Sichevich, O. M. , R. V. Lapunova, A. N. Soboley, Yu. N. Grin and Va. P.
Yarmulek. Sov. Phys. Crystallogr. 30: 627 ( 1985)
Stadelmaier, H.H., G. Schneider and M. Ellner. J. Less-Common Met. 115:
L11(1986)
Stoloff, N.S. and R.G. Davis. Prog. Metal. Phys. 13: 3(1966)
Strnat, K. J. In: E. P. Wohlfarth and K. H. J. Buschow, eds. Ferromagnetic
84 Z. D. Zhang

Materials, Vol. 4. North-Holland, Amsterdam, p. 131(1988)


Sugimoto, S., H. Nakamura, M. Okada and M. Homma. Proc. 12th Int.
Workshop on RE Magnets and their Applications. the University of West
Australia, Nedland, Australia, p. 372( 1992)
Sui, Y.C., Z.D. Zhang, a.F. Xiao, W. Liu, X.G. Zhao, T. ZhaoandY.C.
Chuang. J. Phys. Condo Matt. 8: 11,231(1996)
Sui, Y.C., Z.D. Zhang, W. Liu, a.F. Xiao, X.G. Zhao, T. ZhaoandY.C.
Chuang. J. Magn. Magn. Mater. 172: 285(1997a)
Sui, Y.C., Z.D. Zhang, W. Liu, a.F. Xiao, X.G. Zhao, T. ZhaoandY.C.
Chuang. J. Phys. Condo Matt. 9: 6293(1997b)
Sui, Y.C., Z.D. Zhang, a.F. Xiao, W. Liu, X.G. Zhao, T. ZhaoandY.C.
Chuang. J. Magn. Magn. Mater. 170: L17(1997c)
Sui, Y.C., Z.D. Zhang, a.F. Xiao, W. Liu, X.G. Zhao, T. ZhaoandY.C.
Chuang. J. Phys. Condo Matt. 9: 9985( 1997d)
Sui, Y.C., Z.D. Zhang, a.F. Xiao, W. Liu, T. Zhao, X.G. ZhaoandY.C.
Chuang. J. Alloys and Compounds. 267: 215 ( 1998a)
Sui, Y.C., Z.D. Zhang, W. Liu, a.F. Xiao, X.G. Zhao, T. ZhaoandY.C.
Chuang. J. Phys. Condo Matt. 10: 363(1998b)
Verhoef, R. , F. R. de Boer, Zhi-dong Zhang and K. H. J. Buschow. J. Magn.
Magn. Mater. 75: 319(1988)
Wang, D. , S. H. Liou, P. He, D. J. Sellmyer, G. C. Hadjipanayis and Y.
Zhang. J. Magn. Magn. Mater. 124: 62(1993)
Wu, E., G. H. J. Wantenaar, S. J. Campbell and H. S. Li. J. Phys.
Condens. Matter 5: L457( 1993)
Wu, E. , Heng Zhang and S. J. Campbell. Phys. Status Solidi B 207: 485
(1998)
Xiao, aun-feng, X. K. Sun, Dian-yu Geng, Wei Liu, Zhi-dong Zhang, Tong
Zhao and Y.C. Chuang. J. Magn. Magn. Mater. 140-144: 1093(1995)
Xiao, a.F., Z.D. Zhang, T. Zhao, W. Liu, Y.C. Sui, X.G. ZhaoandD.Y.
Geng. J. Appl. Phys. 82: 6170(1997)
Xiao, a. F. , T. Zhao, Z. D. Zhang, M. H. Yu, X. G. Zhao, W. Liu, D. Y.
Geng, X. K. Sun and F. R. de Boer. J. Magn. Magn. Mater. 184: 330
(1998)
Yang, X., T. Miyazaki, K Takakura and M. Takahashi. J. Magn. Magn.
Mater. 62: 293(1986)
Yang, Y.C., H. Sun. L.S. Kong, J.L. Yang, Y.F. Ding, B.S. Zhang, C.
T. Ye, L. Jin and H. M. Zhou. J. Appl. Phys. 64: 5968(1988)
Yang. Y.C., X.D. Zhang. L.S. Kong, a. Pan, S.L. Ge, J.L. Yang, Y.F.
Ding, B. S. Zhang, C. T. Ye and L. Jin. Solid State Commun. 78: 313
(1991a)
Yang, Y. C. , X. D. Zhang, S. L. Ge, a. Pan, L. S. Kong and H. L. Li. J.
Appl. Phys. 70: 6001( 1991b)
Yang, Y.C., X.D. Zhang, L.S. Kong, a. Pan and S.L. Ge. Appl. Phys.
Lett. 58: 2042(1991c)
Yu, M. H. , Z. D. Zhang, a. F. Xiao, D. Y. Geng. W. Liu and X. G. Zhao.
New Rare-Earth Transition-Metal Intermetallic Compounds and 85

J. Appl. Phys. 88: 4226 (2000)


Zhang, Dan, Zhi-dong Zhang, Y. C. Chuang, Bai-shen Zhang, Ji-lian Yang
and Hong-nian Du. J. Phys. Cond. Matt. 7: 2587 ( 1995)
Zhang, Heng, S. J. Campbell, S. Bulcock, A. V. J. Edge and E. Wu. Physica
B 229: 333(1997)
Zhang, Heng, S. J. Campbell, E. Wu. S. J. Kennedy, H. S. Li, A. J.
Studer, S. Bulcock and A. D. Rae. J. Alloys and Compounds 278: 239
(1998)
Zhang, Heng, S. J. Campbell, H. S. Li and E. Wu. J. Alloys and Compounds
284: 155( 1999)
Zhang, Heng, S. J. Campbell and A. V. J. Edge. J. Phys. Condens. Matter.
12: L 159(2000a)
Zhang, Heng, S. J. Campbell, H. S. Li, M. Hofmann and A. V. J. Edge. J.
Phys. Condens. Matter. 12: 5021(2000b)
Zhang, X.D., Q. Pan, S.Z. Dong, S.L. Ge, Y.C. Yang, J.L. Yang, B.S.
Zhang, Y.F. Ding and C. T. Yeo Acta Physica Sinica 2: 537(1993)
Zhang, Z. D., W. Liu, D. Zhang, X. M. Jin, X. G. Zhao and Q. F. Xiao. J.
Phys. Cond. Matt. 11: 3951( 1999)
Zhang, Z. D. , W. Liu, J. P. Liu and D. J. Sellmyer. J. Phys. D Appl. Phys.
33: R 217(2000)
Zhao, Xin-guo, Zhi-dong Zhang, Qun-feng Xiao, Wei Liu and X. K. Sun. J.
Magn. Magn. Mater. 148: 419( 1995)
Zhao, X.G., Z. D. Zhang, X. K. Sun and W. Liu. J. Magn. Magn. Mater.
208: 231(1999)
Zhong, Xia-ping, R.J. Radwanski, F.R. de Boer, T.H. Jacobs and K.H.J.
Buschow. J. Magn. Magn. Mater. 86: 333(1990)

This work has been supported by the National Natural Science Foundation of China under
projects Nos. 59725103, 59831010. 50071062 and 50331030 the Sciences and Technology
Commission of Shenyang and Liaoning.
3 Magnetic Properties and Interstitial Atom Effects
in the R (Fe, M) 12 Compounds

Jinbo Yang and Yingchang Yang

3. 1 Introduction

Permanent magnets are now widely used in motor, generator,


telecommunication, and control devices, and are considered as indispensable
materials in everyday life. Until now, the intermetallics composed of the rare
earth and 3d elements have been mainly developed as high-performance
permanent magnets, such as SmCos , Sm2 (Co, Fe) 17 and Nd2Fe14 B (Strnat et
aI., 1967; Ojima et aI., 1977; Buschow, 1977; Mishra et aI., 1981; Sagawa
et al. , 1984; Croat et al. , 1984; Hadjipanayis et al. , 1984; Sellmyer et al. ,
1984). As the main, components Sm and Co are particularly expensive, and
therefore it is desirable to use the less costly iron-based compounds in place of
the cobalt-based compounds. Despite the initial promise of the Nd2Fe14 B
magnet, problems associated with its poor temperature stability and corrosion
resistance have ensured that the search for even better permanent magnetic
materials continues unabatedly. Since 1987, worldwide efforts have been
made to investigate the magnetic properties of R(Fel- x Mx ) 12 (M = Ti, V ,Mo,
Cr,Mn,W, AI or Si)(Yang et aI., 1981; Buschow, 1988, 1991; Li and Coey,
1991; Suski, 1996, and references therein). Among them, Sm (Fe, M) 12
seems to be the best potential candidate for permanent magnet applications
(Schultz and Katter, 1991 a). Unfortunately, Sm ( Fe, M) 12 appears to offer less
practical use than Nd2Fe14 B. The Curie temperature is almost the same as that
of Nd2Fe14 B, but its saturation magnetization is lower than the Sm (Fe, M) 12 ,
which leads to the conclusion that the theoretical maximum energy product
(BH)max of Sm(Fe,M)12 is only half that of Nd2Fe14B. In addition, the high
price of Sm metal is also a disadvantage. Studies of the interstitial atom
effects on the magnetic properties of rare-earth intermetallic compounds,
R2Fel7 (C, N) x (Coey and Sun, 1990) and R ( Fe, M) 12 Nx ( M = Ti, V, Mo,
etc.) (Yang et a!., 1990), have recently made great progress in the field of
hard magnetic materials. The interstitially modified compounds with higher N
or C concentration have been found to have excellent intrinsic magnetic
properties; for more detail see the review article (Fujii and Sun, 1995). The
interstitial atoms not only increase both Curie temperature and saturation
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)'2 Compounds 87

magnetization, but also give rise to a remarkable change in the


magnetocrystalline anisotropy of the host compounds. As a result,
R(Fel-x Mx) 12Nx (R = Nd, Pr) emerge as potential candidates for hard magnets
(Yang et al. , 1991a, 1992, 1993a; Wang and Hadjipanayis, 1991; Anagnostou et
aI., 1991; Endoh et aI., 1992; Hu B. P. et al. , 1994). Many efforts have been
made to develop promising interstitial modified compounds into hard magnets
(Yang et aI., 1994; Gong and Hadjipanayis 1993a, 1993b; Tang et aI., 1993;
Wang et aI., 1993a; Mao et aI., 1997; Yang et aI., 1997a). It is also of
great theoretical interest to study the effect of interstitial atoms (H, C, N) on
the magnetic properties of the 1 : 12 compounds.

3. 2 Interstitial Atom Effects and Intermetallic Compounds

3. 2. 1 Crystallographic Structures

R(Fel- x M x ) 12 crystall izes in the tetragonal ThMn12 body-centered structure,


space group 14/mmm, which can be deduced from the CaCus structure. The
ThMn12 structure is illustrated in Fig. 3. 1. The rare-earth atoms occupy the
single 2a sites, while the transition metals Fe and M occupy the three sites:
8i, 8j, and 8f. The binary phase RFe12 does not exist. The ThMn12 structure,
however, can be stabilized by adding a third element M like Ti, V, Mo, W, Mn,
Si,AI,Ta,Nb,Ge, etc. It has been found that the magnetic properties of the
R(Fel-xMx)'2 compounds can be drastically changed by the introduction of
nitrogen and carbon atoms through a gas-phase interstitial modification (GIM)
process (Yang et aI., 1991 a; Hurley and Coey, 1992; Gong and
Hadjipanayis, 1993). Upon nitrogenation or carbruzation, Nor C atoms enter
into the 1 : 12 structure. The typical X-ray diffraction patterns of NdFe,o.s Mo1.5,
NdFelO. sMo1.5 Nx and NdFelO. sMo1.5 C x are presented in Fig. 3.2. All the

R 2a
Fe(M) 0 Si 8Sf OSj
N(C) 0 2b

Figure 3. 1 The unit cell of the ThMn'2 structure.


88 Jinbo Yang and Yingchang Yang

o
M

30 40 SO 60 70 80
28n
Figure 3. 2 The X-ray diffraction patterns at room temperature for
NdFe,o.sMo1.5' NdFelO.sMo1.5 Nx ' and NdFelO.sMo1.5Cx.

compounds are found to crystallize in the ThMn'2-type tetragonal structure (so-


called 1 : 12 phase). The parent compound before nitrogenation or carbonization is
single phase without any magnetic impurities. However, a small amount of the
a-Fe appears in the nitrides or carbides. All the X-ray diffraction peaks of
NdFelo.sMo1.5Nx and NdFelO.sMo1.5Cx shift to the lower angle. This is evidence
for the fact that nitrogen and carbon atoms enter into the interstitial sites and
results in a lattice expansion. Tables 3. 1,3.2 and 3.3 list the corresponding
lattice parameters of RFelO.S M1.5 and theirs nitrides (R=rare-earth, M=Ti, V,
Mo). Since the nitrogen is smaller than that of the rare-earth and iron atoms,
the unit cell dilation implies that the nitrogen atoms should enter into the
interstitial sites rather than the substitutional sites. The ability to absorb
nitrogen or carbon atoms seems to be different from one rare-earth compound
to another, and also shows a dependence on the third element M. This is
reflected by the variation of the expansion rate 5 V/ V (V is the unit cell
volume). Neutron diffraction was used to determine the sites of the interstitial
atoms in YFelO M02 and NdFelO.s Mo1.5 compounds. Figure 3. 3 is the typical
neutron diffraction patterns of NdFelO. sMo1.5 and its nitrides at room
temperature. The neutron data refinement results of YFel0 M02 , NdFe,o.s Mo1.5.
and their nitrides are listed in Table 3. 4. These data demonstrate that the
nitrogen atoms occupy the 2b sites, whereas the iron atoms occupy the 8i,
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)12 Compounds 89

Table 3. 1 Lattice parameters a and c, unit cell volume V, relative change in the unit cell
volume upon nitrogenation S V / V, Curie temperature Tc ' saturation magnetization a., and
anisotropy fields ~o H a of RFe" Ti and RFell TiN, (Yang et aI., 1991a) (The nitrides are
obtained by heating at 770 K for 4 h).
SV/V Tc a.(A rr-r kg-I) ~o Ha(T)
a (A) c (A) v(N)
(%) (K) 1.5K 300 K 1.5 K 300 K
YFellTi 8.503 4.800 347.6 540 138.0 120.1 4.0 2.5
YFell TiN, 8.611 4.821 357.5 3.0 739 167.5 147.4 3.5 1.8
CeFe" Ti 8.531 4.789 348.5 485 132.1 118.5 6.5 2.2
CeFell TiN, 8.623 4.845 360.2 3.3 729 152.8 144.1 4.1 1.9
NdFell Ti 8.574 4.907 360.7 570 149.3 132.8 - 0.9
NdFell TiN, 8.701 4.844 366.7 1.7 740 158.8 145.4 11.5 8.0
SmFe"Ti 8.557 4.800 351.4 600 131. 4 121. 5 14.2 9.3
SmFell TiN, 8.641 4.788 357.5 1.7 742 144.7 138.5 Plane Plane
GdFellTi 8.548 4.800 350.7 610 84.9 80.5 - -

GdFe" TiN, 8.595 4.782 353.3 0.7 660 110.9 102.6 5.5 4.0
TbFell Ti 8.537 4.808 350.4 580 65.5 73.8 - -

TbFell TiN, 8.581 4.798 353.3 0.8 630 83.4 88.0 14.0 11. 1
DyFell Ti 8.521 4.799 348.4 560 66.4 74.6 - -

DyFell TiN, 8.570 4.798 351.2 0.8 620 91.0 92.3 14.0 12.5
HoFell Ti 8.506 4.799 347.2 540 64.7 82.0 - -
HoFell TiN, 8.561 4.787 350.8 1.0 618 85.6 96.3 11.0 9.0
ErFell Ti 8.495 4.795 346.0 530 66.7 84.5 8.1 3.6
ErFell TiN, 8.548 4.792 350.1 1.2 610 101. 9 103.3 Cone 3.5

Table 3. 2 Lattice parameters a and c, unit cell volume V, relative change in the unit cell
volume upon nitrogenation SV/ V, Curie temperature Tc ' saturation magnetization a., and
anisotropy fields ~oHa of RFelO.sMo1.5 and RFelO.sMo1.5N, (The nitrides are obtained by
heating at 770 K for 2 h). Continued on page 90.
SV/V Tc a. (A' rr-r' kg- 1 ) ~o Ha(T)
a (A) c (A) v(N)
(%) (K) 1.5 K 300 K 1.5K 300 K
YFe,o.sMo1.5 8.543 4.793 349.8 453 129.5 99.4 3.0 1.5
YFelOS Mol .SN, 8.659 4.799 359.6 2.9 642 142.1 124.5 2.0 1.0
CeFel0 sMo1.5 8.550 4.773 348.9 379 111. 4 73.0 2.0 1.0
CeFeIO. sMo1.5 N, 8.641 4.829 360.6 3.3 604 131. 6 110.0 1.0 0.5
PrFelO. sMo1.5 8.607 4.796 355.3 455 134.4 106.7 Plane Plane
PrFeIO.S Mo1.5 N, 8.672 4.869 366.1 3.0 641 140.5 119.2 16.3 11.0
NdFelO. sMo1.5 8.590 4.791 353.5 470 140.6 101. 5 Cone 0.6
NdFelO.S Mo1.5 N, 8.693 4.870 368.0 4.0 665 144.5 120.1 18.1 13.3
SmFelO. sMo1.5 8.587 4.789 353.1 491 121. 3 100.2 14.2 9.3
SmFelO.S Mol.S N, 8.623 4.817 358.2 1.4 620 130.3 112.2 Plane Plane
90 Jinbo Yang and Yingchang Yang

Table 3. 2 Continued from page 89.


5V/V Tc as CA' rrf kg-') Jio H a (T)
a (A) c (A) v(N)
C%) CK) 1.5K 300 K 1.5K 300 K
GdFelo. sMo,. s 8.580 4.801 353.4 535 89.2 75.7 5.0 3.0
GdFelo sMo1.5 Nx 8.651 4.826 361.2 2.2 702 110.8 99.9 1.5 1.0
TbFe,0 sMo1.5 8.555 4.794 350.9 467 53.4 53.2 - -

TbFe,o.s Mo1.5 Nx 8.620 4.795 356.3 1.5 650 81. 0 76.4 18.1 15.2
DyFe,o.s Mo1.5 8.542 4.788 349.4 457 53.4 60.7 - -
DyFelo. sMol. sNx 8.614 4.808 356.8 2.2 646 70.6 72.3 15.6 12.6
HoFelO. sMo1.5 8.528 4.782 347.8 445 65.8 73.8 3.0 1.0
HoFelO.sMo,.sN x 8.643 4.807 359.1 3.2 641 81.5 85.2 13.0 7.0
ErFelO.S Mo1.5 8.520 4.784 347.3 430 67.4 76.2 - 1.0
ErFelO.S Mo1.5 Nx 8.574 4.786 351.8 1.3 632 80.6 90.1 Cone 1.0

Table 3. 3 Lattice parameters a and c. unit cell volume V. relative change in the unit cell
volume upon nitrogenation 5 V/ V. Curie temperature Tc. saturation magnetization a..
anisotropy fields JicHa of RFe,0.S V1.5 and RFe,o.s V1.5 Nx(The nitrides are obtained by heating at
and

790 K for 5 h)
Tc a.CArrfkg-') Jio H a CT)
a CA) c CA) VCN) 5V/V
C%) CK) 1.5K 300 K 1.5K 300 K
YFe,o.s V1.5 8.488 4.770 343.6 582 136.9 129.1 5.6 2.2
YFe,o.s V1.5Nx 8.609 4.787 354.7 3.2 793 154.1 145.8 4.2 1.3
CeFe,o.s V1.5 8.519 4.758 345.3 508 125.6 114.1 5.0 2.8
CeFelO. sV1.5 Nx 8.576 4.802 353.2 2.3 758 142.4 136.4 4.0 1.7
PrFe,o.s V1.5 8.556 4.773 349.4 625 131. 4 121. 4 Plane Plane
PrFelO.S V1.5 Nx 8.638 4.837 360.9 3.3 820 157.5 142.8 15.3 10.8
NdFe,o.s V1.5 8.5598 4.775 349.8 620 143.9 131. 0 - 1.0
NdFelO. sV1.5 Nx 8.607 4.907 363.5 3.9 860 148.8 139.1 18.6 11.0
SmFe,o.s V1.5 8.540 4.774 348.2 650 128.6 115.4 13.8 8.2
SmFe,o.s V1.5 Nx 8.609 4.808 356.3 2.3 802 140.5 129.8 Plane Plane
GdFelO.s V1.5 8.521 4.776 346.8 658 79.6 73.6 6.1 3.2
GdFelO.S V1.5 Nx 8.575 4.779 352.9 1.8 820 98.0 92.2 5.0 2.9
TbFe,o.s V1.5 8.494 4.772 344.3 618 79.8 87.3 - 0.8
TbFelO. sV1.5 Nx 8.576 4.790 352.3 2.3 819 105.7 101.2 21. 5 12.4
DyFelO.S V1.5 8.483 4.770 343.3 590 68.6 76.1 - 2.5
DyFe,o.s V1.5N x 8.614 4.803 356.3 3.8 819 98.1 88.9 20.0 13.0
HoFe,o.s V1.5 8.478 4.765 342.5 575 69.3 81.0 6.0 2.8
HoFe,o.s V1.5Nx 8.625 4.780 355.5 3.8 800 99.0 104.3 13.5 8.0
ErFe,o.s V1.5 8.470 4.765 341.8 553 66.0 83.4 - 2.6
ErFe,o.s V1.5 Nx 8.567 4.768 349.9 2.4 784 85.9 101. 2 Cone 2.2
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)12 Compounds 91

8 NdFelO.5Mo1.5 Yobs
- - Yeal
- . - Yobs-Yeal

20 40 60 80 100
2en
Yobs
- - Yeal
- . - Yobs-Yeal

C-4
~
o
U 0

20 40 60 80 100
2en
Figure 3. 3 The neutron diffraction patterns at room temperature for
NdFelO.5 Mo1.5 and its nitride.

Table3.4 Crystallographic sites, occupations of the atoms, coordinates x, y, z, and M


magnetic moments (/.Is) on the different sites of YFe,O M02 , YFelO M02 NIl' NdFelO.5 Mo1.5 and
NdFelO.5 Mo1.5 N1.2. Continued on page 92.
______1 occupations I x I y I z ~
YFelOM02 a=b=8.576(1)A c=4.811(1)A V=353.8N To =360 K
Y(2a) 1 0 0 0
Fe,Mo(8D 0.44,0.56 O. 358( 1) 0 0 1. 3(3)
Fe(8j) 1 O. 279( 1) 0.5 0 1.3(3)
Fe(8t) 1 0.25 0.25 0.25 1.3(3)
YFe,O M02 N, .1 a=b=8.694(1)A c=4.816(1)A V=364.oN To =460 K
Y(2a) 1.0 0 0 0
Fe(8i), Mo(8D 0.482,0.522 0.361(1) 0 0 1.6(3)
Fe(8j) 1.0 0.282(1) 0.5 0 1.6(3)
Fe(8t) 1.0 0.25 0.25 0.25 1. 6(3)
N(2b) 0.99 0 0 0.5
NdFelO.5 Mo1.5 a = b =8. 585(2) A c=4.788(1)A V=352.9N To =461 K
Nd(2a) 1 0 0 0 1. 6(4)
Fe,Mo(8i) 0.60,0.40 0.3592(2) 0 0 1. 6(4)
92 Jinbo Yang and Yingchang Yang

Table 3. 4 Continued from page 91 .


occupations x y z M(J..Ie)
Fe(8j) 1 0.2745(2) 0.5 0 1. 6(4)
Fe(8f) 1 0.25 0.25 0.25 1.6(4)
NdFelo 5 Mo1.5 N1.2 a = b =8. 663(1)A c = 4. 8653(7)A V=365.1N T c =650 K
Nd(2a) 1 0 0 0 1. 6(2)
Fe,Mo(8D 0.6,0.4 0.3592(5) 0 0 1. 7(2)
Fe(8j) 1 0.2740(5) 0.5 0 1. 7(2)
Fe(8f) 1 0.25 0.25 0.25 1. 7 (2)
N(2b) 0.99 0 0 0.5

8j and 8f, and Mo occupies the 8i sites. A number of neutron diffraction studies
on YFe" TiN x ' YFelO V2 Nx and YFel0 M02 Nx or YFell MoN x have been made
previously (Yang et aI., 1991 d; Sun et aI., 1993a, 1993b; Yelon and
Hajipanayis, 1992; Yang et aI., 1997b). All studies present the same conclusion
that the nitrogen atoms occupy the octahedral 2b interstitial sites. Accordingly,
full occupation of the 2b sites in the case of R (Fe, M) 12 N (C) x leads to a
composition of R (Fe, M) 12 N (C) 1.0' However, the experimentally estimated
nitrogen or carbon contents ranges from 0 to a value as large as 2.7 (Yang et
aI., 1998a, 2000a; Goto et aI., 1995; Oleinek et al. , 1998), which is much
higher than the maximum value of one nitrogen (carbon) atom per formula unit
(1. u.). Such an excess of the interstitial atoms can be understood assuming
that the carbon and nitrogen atoms are additionally trapped on the surfaces
and/or in the grain boundaries (Yang et aI., 1998c). A similar excessive
absorption of the nitrogen atoms has also been observed during the
nitrogenation of R2Fe17 (Wei et aI., 1993; Iriyama et al. ,1992). Besides the
GIM process (Hurley et al. , 1992; Li. et aI., 1993; Oi et al. , 1992; Oleinek
et al., 1999a), carbon atoms can also be introduced into the 1 : 12 crystal
lattice by: CDarc melting or induction melting of the constituent elements (Yang et
al. , 1993b; Hu. et aI., 1996; Mao et aI., 1998a ), (2) synthesis from heavy
hydrogen-carbon compounds (Fruchart et aI., 1994), @ solid-solid phase
reaction (Skolozdra et aI., 1995; Izumi et aI., 1996), and @ mechanical
alloying of the starting elements (Schultz et aI., 1991b). On the other hand,
neutron diffraction studies have shown that the carbon atoms may enter the
substitutional sites in the carbides produced by arc melting (Hu et aI., 1996;
Mao et aI., 1998a). For hydrogen, it has been known as interstitial in the
rare-earth transition metal intermetall ics for a long time, and a very small
expansion (0.2 % - 1. 0 %) of the unit cell volume is observed in the 1 : 12
compounds (Zhang et aI., 1990; Obbade et aI., 1988; Tomey et aI., 1996;
Mao et aI., 1998b). The neutron diffraction results indicate that the H atoms
occupy the interstitial 2b sites. In Table 3. 5, we summarize the structural and
magnetic properties of some carbides and hydrides.
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M),2 Compounds 93

Table 3. 5 Lattice parameters a, C , unit cell volume V, relative change in the unit cell
volume 5V/ V , Curie temperature To' saturation magnetization as, and anisotropy fields
IJo H. of some carbides and hydrides.
5V/V To asCA-ni-kg-') IJo H. (T)
a (A) c CA) VCN)
C%) CK) 10 K 293 K 293 K
NdFell Ti 8.590 4.797 354.0 549 148.8 108.1 0.9
NdFell TiC 14 8.607 4.864 359.6 1.6 650 152.9 111.1 4.0
NdFell TiC 2 . 1 8.609 4.919 364.6 3.0 675 157.7 120.0 5.5
NdFelO. 75 V1.25 8.541 4.754 346.8 607 139.9 110.1 0.9
NdFe,O.75 V1.25C1.5 8.550 4.848 354.4 2.2 670 141. 9 123.0 4.7
NdFelO. 75 V1.25 C 2 3 8.561 4.878 357.5 3.1 692 148.8 127.9 5.4
NdFe,o 5Mo1.5 8.564 4.798 351.9 470 148.9 97.2 0.5
NdFe,O.5 Mo1.5 C1. 4 8.594 4.852 358.4 1.8 579 151. 7 102.1 6.4
NdFelO.5 Mo1.5 C z.z 8.603 4.875 360.8 2.5 602 151. 9 109.1 7.2
YFe,o 5Mo1.5
(1)
8.540 4.791 349.4 433 78.7 0.9
(1)
YFe,O.5 Mo1.5 Co.z 8.546 4.792 350.0 0.2 460 87.4 1.0
(1)
YFe,O.5 Mo1.5 Co 4 8.560 4.794 351.3 0.4 480 95.9 0.8
(1)
YFelO.5 Mo1.5 CO.6 8.584 4.800 353.7 1.1 488 101. 7 0.7
NdFe,O.5 M01.5CO.4
(1)
8.613 4.803 356.3 1.8 452 110.4 0.9
(1)
DyFelO.5 Mo1.5 8.543 4.793 349.8 457 52.6 1.7
DyFelO.5 Mo,. 5Co. 6
(1)
8.560 4.794 351.3 0.4 490 69.0 3.5
YFelOMoz 8.563 4.810 352.7 410 105.2 65.8 1.9
YFelOMazHx 8.592 4.817 355.6 0.8 440 108.9 77.1 3.2
YFelOMozCx 8.603 4.816 356.4 1.0 472 110.5 78.2 0.5
YFe"Ti 8.522 4.797 348.4 530 140.1 120.3 2.1
YFell TiH x 8.546 4.800 350.6 0.6 580 146.6 132.9 2.5
YFe"TiCx 8.561 4.812 352.7 1.2 655 152.7 134.5 1.2
YFelO.5 V1.5 8.488 4.770 343.7 580 126.5 114.3 2.2
YFelO.5 V1.5 Hx 8.503 4.778 345.5 0.52 630 131. 3 121. 1 2.6
( 1) Compounds are prepared by arc-melting method.

3_ 2_ 2 Curie Temperature and Saturation Magnetization


The magnetic properties of the R ( Fe, M) 12, their nitrides, carbides and
hydrides are also listed in Tables 3. 1- 3.3 and 3.5. The average increase in
the Curie temperature (T c) is about 200 K for the nitrides, 100 K for the
carbides and 50 K for the hydrides. Since yttrium is nonmagnetic, T c of the
yttrium compound is determined mainly by the Fe-Fe interactions. The fact that
the T c of Y(Fe, M) 12 Nx is increased about 200 K demonstrates that the Fe-Fe
94 Jinbo Yang and Yingchang Yang

interaction is enhanced by introduction of the nitrogen atoms into the parent


compounds. The Curie temperatures of RFel0.5 M1.5 and RFelO.5 M1.5 Nx (M = V,
Mo, Tj) are plotted in Fig. 3. 4. Unexpectedly, there is different behavior in
the variation of T c across the series of nitrides. In contrast with the parent

: RFelO.SM01.5 0: RFelO.sM0I.5N
: RFelO.5VI.5 0: RFelO.sVI.5N
900 ~: RFellTi 6: RFel1 TiN.

800

700

600

500

400

Ce Pr Nd Sm Eu Gd Tb Oy Ho Er Y

Figure 3.4 The Curie temperatures as a function of rare-earth R for the R(Fe, M) 12 and
R(Fe,M) 12N. (M= Ti, V ,Mo).

compounds, in which T c obeys the (gJ - 1)2 J (J + 1) law (gJ is Lande g


factor,J is angular momentum quantum number), the familiar variation of T c
across the series with the maximum at Gd is not observed in the nitrides for Ti
and V series. These results suggest that T c also depends on the nitrogen or
carbon content. The variation of 5 V/ V across the series indicates that the
nitrogen content is different from one rare-earth compound to another.
Generally, T c of the R-Fe compounds is determined by the Fe-Fe, R-Fe and
R-R interactions. The R-R interactions can be ignored in the R-Fe compounds.
Assuming the effective iron spin S = 1, and neglecting the R-R interaction, the Curie
temperature Tc can be expressed in the molecular field model (Coey et aI., 1989),

Tc = ~ (T Fe + Jne + 4nFe) }
TFe = nFe-Fe N Fe [4 x S(S + 1)p~/3kBJ (3.1)
TR- Fe = nR-Fe Iyl (N R N Fe )112{2g[S(S + 1)J(J + 1)Jl/2p~/3kB}

where N Fe and N R are the number of the atoms per un it volume, n Fe-Fe and n R-Fe
are the molecular field exchange coefficients, y = 2 (g - 1) / g, where 9 is the
Lande g factor and J is the rare-earth angular momentum quantum number.
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)'2 Compounds 95

The average values of the nFe-Fe and nR.Fe for the RFe,o.5 T1.5 and RFelO.5 T1.5 N.
(T = Ti, Mo, V) series compounds were deduced from the T c data of the
heavy rare-earth compounds using Eq. (3. 1 ), ignoring the lanthanide
contraction effect. The values obtained are listed in Table 3.6. It is found that
the Fe-Fe exchange interaction increases about 40 % due to the introduction of
the nitrogen, however the R-Fe interaction decreases due to the formation of
the chemical bonding between the Rand N atoms. This effect also was
observed in the RFell TiC. carbides (Hurley et aI., 1992).

Table 3. 6 Molecular field exchange coefficients ( nFe-Fe ) and (nR-Fe) of the R(Fe, M)'2 and
R(Fe,M)'2Nx compounds.
(nFe-Fe) (T 1.u. IlJa) ( nR-Fe) (T 1. u. I lJa)
RFe"Ti 25.9 14.8
RFell TiN x 30.3 11.6
NdFe,o.5 Mo1.5 21.3 15.4
RFelO.5 Mo1.5 Nx 32.6 14.2
RFelO.5 V1.5 28.1 17.6
RFe,O.5 V1.5 Nx 41.9 11. 1

In comparison with of the parent compounds, it is evident that a large


increase in the spontaneous magnetization is achieved upon nitrogenation.
Since the 4f electrons of the rare-earth atoms are well localized, in principle,
the magnetic moments of the rare-earth ions are close to gJ J Pe. Accordingly,
the effect of increasing the spontaneous magnetization is related to a
modification of the 3d band. From the experimental data for the yttrium
compounds where the 4f magnetic moment is absent, it can be derived that the
introduction of the nitrogen atom's leads to an increase in the Fe moment. The
iron magnetic moments deduced from the experimental values of YFe,o.5 T1.5 N.
(T = V, Mo) are about 2.0 Peat 1.5 K and 1.8 Peat room temperature. The
increases of the atomic magnetic moment of the Fe atom's are approximately
12 % - 24 % at 1. 5 K and 14 % - 28 % at 300 K compared to the values before
nitrogenation. The interstitial atoms modification with carbon atoms leads to a
smaller increase of magnetization compared to the increase upon
nitrogenation. The increase of the Fe magnetic moments in the carbides is
about 15 % at 300 K and 5 % at 10K compared to the parent compounds. An
insight into these effects will be gained from the band structure calculations
given in the following Section 3.3.2.

3.2.3 Magnetocrystalline Anisotropy


On the basis of the experimental results, the magnetocrystalline anisotropy of
R(Fe, M),2N. nitrides can be classified as follows:
96 Jinbo Yang and Yingchang Yang

C1) The c-axis is the easy magnetization direction from 0 K to the Curie
temperature where R represents Pr, Nd, Gd, Tb, Dy and Ho;
(2) The easy magnetization direction lies in the basal plane of 8m;
(3) A spin reorientation occurs at low temperature, e. g., ErFell TiN x
nitrides.
Carbon atoms have almost the same effect on the anisotropy as does
nitrogen in the 1 : 12 compounds using the GIM process. However, the carbon
atoms have little effect on the anisotropy of the arc-melted carbides CYang
et al., 1993b; Mao et al., 1998a). This is due to the different site
occupations of the carbon atoms in this case, and also to the available
maximum carbon content of less than O. 5 atoms/f. u. in the arc-melted
carbides. A first order magnetization process CFOMP) was detected in the
NdCFe,M) 12 C x C M=Ti, V, Mo)CYang et aI., 2000a). The magnetization
curves of NdFell TiC2. 1 parallel and perpendicular to the alignment direction at
different temperatures are plotted in Fig. 3.5. An anomalous increase of the
magnetization with field is found when the external field is applied
perpendicularly to the alignment direction. This phenomenon may be regarded
as a FOMP, which indicates a drastic change in the crystal field CCF) acting
on the rare earth. No FOMP was found in the nitrides of the NdCFe, M),2 Nx
CM = Ti, V, Mo) with a field up to 50 T CMUlier et al. , 2001). The calculated
CF parameters show that the higher-order CF parameters of the Nd ions and
the thermal evolution of the related anisotropy constants play a key role in
producing the FOMPs of NdCFe,M),2CxCYang et aI., 2000a).
1.5
II

..L

f\
}\
0: 10K
A 6: lOOK
-- ---
5
.
,J ...

10
"
'.
"",".
15
H (T)
150K
0:
0: 200K
o 5 10 15
j./{)H (T)

Figure 3.5 Magnetization curves of a magnetically-oriented-powder sample of NdFell Tic... 1 at


different temperatures. Field and measured magnetization perpendicular(closed symbols)
and parallel (open symbols) to the alignment direction. The FOMP field HI is given by the
maximum of the derivative dJ/ dH (inset).
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)12 Compounds 97

It is reasonable to assume that the magnetocrystalline anisotropy comes


from both the R and Fe sublattices. Since gadolinium has no orbital moment
and yttrium is nonmagnetic, the anisotropy of GdCFe, M)12 N. and YCFe, M)12 N.
comes from the Fe sublattice only. The X-ray and magnetic measurements
results indicate GdCFe,M) 12 N. and YCFe,M)12N. have an easy axis, and no
large difference is found in the magnitude of the anisotropy fields before and
after nitrogenation. Therefore, it can be concluded that the interstitial nitrogen
has little effect on the magnetocrystalline anisotropy for the Fe sublattice.
However, a substantial change in magnetocrystalline anisotropy of the R
sublattice is manifested upon nitrogenation. In contrast to Sm CFe, M) 12 ,
SmCFe,M)12NCC). has an easy plane, whereas Nd CFe, M)'2 N CC).,
PrCFe,M)12NCC)., TbCFe,M)12N., DyCFe,M)12N., and HoCFe,M)12N. have
an easy axis from 0 K to room temperature. The characteristics of the
anisotropy in these compounds can be understood by the effect of the nitrogen
Ccarbon) atoms on the crystal-field interactions.

3.2.4 Hyperfine Interactions

Mossbauer spectroscopy is an effective experimental technique for obtaining


information on the interstitial modified effects CLi et aI., 1992, 1993; Sun
et al., 1993c; Oi et al., 1992; Wang et al., 1993b). Representative 57 Fe
Mossbauer spectra of RFelO.5 Mo1.5 and RFelO.5 Mo1.5 N. measured at 77 K are
shown in Fig. 3.6. The intensities of the subspectra were calculated according
to a previous model by Sinnemann et al. CSinnemann et al. , 1989a, 1989b) for
the compounds with the ThMn 12 structure. There are 11 subspectra for fitting
the entire Mossbauer spectrum. For the hyperfine fields B hf at the three sites,
we took the order B hf C8 j) > B hf C8j) > B hf C8t) for the parent compounds,
based on most researchers' analyses and opinions. However, according to the
neutron diffraction data CSun et aI., 1993b) and band calculations CIshida
et aI., 1994; Yang et aI., 1997b, 1998d), the magnetic moments on the
different Fe sites are in the order of 11 C8D > 11 C8t)> 11 C8j) in the interstitial
compounds. Therefore, we took the order of B hf C8D > B hf C8j)> B hf C8t) in
the nitrides, which was different from the previous studiesC Li et al. , 1992; Oi
et aI., 1992, 1994). The fitted hyperfine parameters are listed in Table 3.7.
For RFelO.5Mo1.5' Bit of the 8i sites is about 5 and 10 T more than those on the 8j
sites and 8f sites, respectively. For RFelO.5 Mo1.5 N. as compared to RFelO.5 Mo1.5 ,
the average B It increases by about 4 and 2 T at room temperature and 77 K,
respectively. Upon nitrogenation, Bhfincreases by 5-10 T and 5-12 T for the
8i and 8f sites at room temperature, respectively. The value of B It decreases
2 - 5 T for the 8j sites at room temperature, which is due to the strong
hybridization between the FeC8j) and its neighbor nitrogen atoms CYang, 1997a).
98 Jinbo Yang and Yingchang Yang

1.00 ~~~~~~tii
0.99
0.98
0.97
0.96
1.00

0.99
0.98

0.97
Relative Velocity (mm/s)

c c
0 0
'(ij '(ij

'S'" S'"
'" '"
!:l !:l
<l) <l)
.::(;j .::
(;j
~ ~
~ ~

Er
0.97

--{j -4 -2 0 2 4 6 --{j -4 -2 0 2 4 6
Relative Velocity (mm/s) Relative Velocity (mm/s)
(a) (b)

Figure 3.6 The 57 Fe MOssbauer spectra at 77 K for RFelo.5 Mo1.5 (a) and RFelo5 MoI.5Nx (b) ,
Magnetic Properties and Interstitial Atom Effects in the R( Fe, M) '2 Compounds 99

Table 3. 7 Hyperfine parameters: hyperfine field B"" average hypefine field (B",) average
isomer shift (IS) and average quadrupole splitting (as) for RFeIO.5M01.5 and RFel0.5 Mo1.5 N.
at 77 and 293 K.
T=293 K

B",(T) (B", ) ( IS) (as)


(T) (mm/s) (mm/s)
8i 8j 8f
YFelO.5 Mo1.5 23.2 18.5 14.1 14.6 -0.314 0.052
YFelO.5 Mo1.5 N. 28.5 14.9 21.4 20.3 -0.253 - 0.035
CeFelO.5 Mo1.5 21.9 17.9 13.3 18.1 -0.236 -0.026
CeFelO.5 MO I.5N. 31.9 18.3 26.2 24.3 -0.181 0.023
PrFeIO.5 Mo1.5 24.6 19.3 13.9 18.3 -0.270 0.007
PrFelO 5Mo1.5 N. 28.7 16.1 22.4 21.2 -0.202 -0.021
NdFeIO.5 Mo1.5 23.9 18.9 13.4 18.9 -0.301 0.001
NdFeIO.5 Mo u N. 29.5 16.7 22.9 21.8 -0.259 0.012
SmFe,O.5 Mo1.5 27.4 21.2 15.2 20.1 -0.262 0.050
SmFe,O.5 Mo u N. 32.4 16.9 24.4 23.1 -0.227 -0.008
GdFelO.5 Mo u 26.0 20.5 14.4 19.3 -0.306 -0.025
GdFeIO.5 Mo u N. 29.5 16.0 22.5 21.4 -0.257 0.046
TbFe,O.5 Mo u 25.2 19.2 12.8 17.9 -0.260 0.057
Tb Fe IO.5 MOI.5 N 25.6 13.6 19.8 18.6 -0.220 0.014
DyFelo.5 Mo u 23.4 18.1 12.4 17.0 -0.300 0.028
DyFelo.5 Mo u N. 30.5 14.6 22.3 21.0 -0.237 0.039
ErFeIO.5 Mo u 21.9 17.9 13.1 14.7 -0.309 0.016
ErFeIO.5 Mol. 5N. 26.5 15.1 18.4 18.1 -0.280 0.046
T=77 K
YFelO.5 Mo u 29.6 24.0 17.3 22.5 -0.189 0.016
YFeIO.5 Mo1.5 N. 32.4 20.2 26.5 25.3 -0.144 -0.013
CeFeIO.5 Mo u 29.7 25.1 19.2 23.7 - 0.186 0.031
CeFeIO.5 Mo1.5 N. 34.6 22.0 29.9 27.8 -0.155 -0.008
PrFe,O.5 Mo u 31. 5 25.1 18.4 23.8 -0.185 -0.015
PrFelO.5 Mo1.5 N. 32.3 19.6 26.6 25.0 -0.128 0.068
NdFelO.5 Mo u 30.7 24.1 17.5 22.9 - 0.188 -0.072
NdFeIO.5 Mo u N. 32.1 19.8 26.3 25.0 -0.147 0.014
SmFel0.5 Mo u 30.7 25.2 18.7 23.8 -0.191 0.023
SmFel0.5 MOuN. 34.2 21.0 28.1 26.6 - 0.152 -0.009
GdFeIO.5 Mo u 30.6 24.9 19.0 23.7 -0.198 0.016
GdFeIO.5 Mo u N. 32.1 19.5 25.3 26.0 -0.153 0.017
TbFel0.5 Mo u 30.7 25.0 18.7 23.2 -0.198 0.016
TbFe IO.5Mo l .5N. 32.6 20.1 26.5 25.3 -0.167 0.019
DyFelo.5 Mo u 29.2 23.3 17.0 22.0 -0.300 0.028
DyFeIO.5Mol.5N. 33.5 20.0 27.4 25.8 -0.161 0.006
ErFel0.5 Mo u 28.5 22.7 15.9 21.2 -0.168 0.032
ErF e IO.5 Mo l.5 N 30.9 17.1 24.3 22.8 -0.148 0.008
100 Jinbo Yang and Yingchang Yang

The average isomer shift CIS) of RFelO.5 M0 15 increases after nitrogenation for
the entire series. There are two factors, which are largely responsible for the
change of the hyperfine fields and the isomer shifts. One is the magnetovolume
effect caused by expansion of the unit cell volume upon nitrogenation; the other
is the chemical bonding effect mainly due to charge transfer and hybridization
between the interstitial atoms and the neighboring Fe (Yang et aI., 1997b,
1998d). Attempts have been made to separate the volume effect and the
chemical effect (Li et aI., 1992; Yang et aI., 1998c), the results of which
indicates that both the expansion of the unit cell volume and the chemical effect
decrease the s charge density at the nucleus sites, and thus increase the
isomer shift.
The dependence of the hyperfine interaction on the nitrogen content has
also been observed in NdFelO.5V15Nx nitrides (Goto et aI., 1995; Yang et aI.,
1998c). Figure 3.7 shows the 57Fe M6ssbauer spectra of NdFelO.5Mo15Nx
measured at room temperature. The values of B hf and the Is are plotted in
Fig. 3.8. The average hyperfine field (B hf) and the hyperfine fields B hf at
different Fe sites monotonically increase with nitrogen content x. The isomer
shift of all compounds also steadily increases with the nitrogen content.

0.99
0.98

0.97
1.00
0.99
0.98
l=: 0.97
.~ 1.00
en
s 0.99
en
0.98
!:J
.~ 0.97
0;
~ 1.00
0.99
0.98
0.97
1.00
0.99
0.98
0.97

-8 -4 0 4 8
Relative velocity (mm/s)
Figure 3. 7 The 57Fe Mossbauer spectra at room temperature for NdFe105 Mo1.5 Nx .
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)'2 Compounds 101

30

E 25
"0
v<;:;
<l)
c
't
<l)

~ 20
::r:

15
0 0.2 0.4 0.6 0.8 1.0
Nitrogen content,x

-D.15

~
E
S -D.20
:
:.a
'"~
E
0
.~ -D.25
OJ)

~
-<
-D.30
0 0.2 0.4 0.6 0.8 1.0
Nitrogen content, x

Figure 3. 8 The dependence of the hyperfine fields and isomer shifts


on the nitrogen content of NdFelO.5 Mo1.5 Nx .

3.3 Origin of the Interstitial Atom Effects

3.3. 1 Crystal Field Effect


In order to analyze the origin of the magnetocrystalline anisotropy, total
magnetic anisotropy energy E has to be discussed. This energy consists of the
contribution of the 3d electrons from the Fe sUblattice and the contribution of
the rare-earth (R) sublattice . As usual, the contribution of the 3d system to
the magnetization as well as to the anisotropy is assumed to be given by that
102 Jinbo Yang and Yingchang Yang

of the corresponding Y compounds and is determined experimentally.


Since the aspherical orbital wavefunction of the 4f electrons interacts
strongly with the crystal field, the R ions make a great contribution to the
anisotropy of R (Fe, M) 12 N (C) x' The anisotropy energy with a tetragonal
symmetry can be phenomenologically expressed as
(3.2)

The rare-earth part of the magnetic anisotropy E 4f is transferred into the 3d


system, mainly determining the magnetization J, by an effective 3d-4f
exchange interaction:
(3.3)

where H m is the molecular field and J is the total angular-momentum operator


of the R ion. H m is mainly determined by the magnetization and is therefore
temperature dependent. The magnetic anisotropy of the R sublattice is due to
strong crystalline electric fields (CEF) characterized by the CF parameter B rm
(crystal field coefficients) acting on the aspherical 4f wave functions of the R
ions in the nitrided or carburized rare-earth compounds. The R ions in the
ThMnlrtype structure on the 2a sites have 14/mmm point symmetry. Thus the
CEF Hamiltonian Hcf is given by
(3.4)

with Bnm=en<r~f>Anm' en are the Stevens coefficients,<r~f> are expectation


values of r~f for the 4f shell of the considered R ion, A rm are the crystal-field
coefficients, and 0 nm are the Stevens operators. As the 4f-electrons of the R
ions are well localized, the contribution of the R sublattice to the
magnetocrystalline anisotropy can be calculated using the single-ion model.
Thus, the total perturbation Hamiltonian for the R ions is
(3.5)

which takes into account the CEF and the coupling between the 3d and 4f
electrons. Then the contribution of the 4f electrons to the anisotropy is given by

(3.6)

where N is the density of Nd ions, and E; is the eigenvalue of H 4f obtained by


solving the secular equation
(3.7)

where L is the orbital angular momentum, S is spin, and J is the total angular
momentum and M', M are magnetic quantum numbers. Comparing the result
following from Eq. (3.2) with Eq. (3.6), the contribution of the R sublattice
to the anisotropy constants K 1 ( T), K 2 ( T), and K 3 (T) were obtained.
Some calculated results of both parent RFel0.5 M1.5 compounds, nitrides
and carbides are listed in Table 3.8. The results show that the contribution of
neighboring nitrogen (carbon) ions to the second-order crystal-field coefficient
Magnetic Properties and Interstitial Atom Effects in the R( Fe, M) 12 Compounds 103

A 20 is positive and large, while the contribution of the neighboring rare-earth


ions to A 20 is negative, but small, and thus the second-order-crystal coefficient
becomes positive in contrast with that of the parent compounds. Usually, the
second-order crystal-field interaction is dominant. Accordingly, for the nitrides
( carbides), the rare-earth ions with negative second-order Stevens
coefficients a J (Pr ,Nd, Ho, Tb, and Oy) may have easy axes, whereas those
with a positive a J ' such as Sm, Er and Tm, are expected to have an easy
plane(Yang et aI., 1991c; Li and Cadogan, 1992).

Thble 3.8 Orystal-field coefficients Arm of the rare earth sublattioe in some 1 : 12 compounds.
A 20 A. o A" A60 A 66
(Kao 2 ) (Kao') (Kao' ) (Kao 6 ) ( Ka o 6 )

NdFell Ti -30 -14.4 4.69 -0.026 -0.024


NdFell TiN 88.0 21.5 0.839 3.84 0.173
NdFe,05 V u -112 -0.51 5.22 -0.025 -0.037
NdFelO.5 V u N 897 12.82 1. 51 2.13 -0.013
NdFe,O.5 VuO 64.2 31.87 0.64 -1.54 -0.038
PrFe'O.5 V u -25.3 -6.69 -0.035 3.58 0.022
PrFelO 5V 1.5 N 117 40.2 3.09 -1.34 0.037
PrFelO 5Mo1.5 -883 -4.24 10.48 -0.032 -0.018
PrFelO.5 Mo1.5 N 9293 112.73 -5.51 2.73 0.075

3.3.2 Band Structures


Electronic structure calculation is a useful tool to study the effect of the
interstitial effect, especially the change of magnetic properties(Jaswal et al. ,
1991; Jaswal, 1993; Sakuma, 1992; Li and Coey, 1992; Fernando et al. ,
1993; Hu et al. , 1994; Asano et al., 1993; Ishida et al., 1994). It is generally
assumed that the following two effects are responsible for the modification of
the properties after interstitial doping: (Dthe pure magneto-volume effect due
to the increase of the crystal unit cell volume; (g)the chemical effect due to the
charge transfer and hybridization between the interstitial atoms and the
neighboring atoms. Experimentally, these two effects can hardly be
separated. Ab-initio electronic theory is a useful method for solving this
problem. The linear muffin-tin orbital (LMTO) method (Anderson, 1975) in
atomic-sphere approximation (ASA) has been employed to performed a semi-
relativistic band calculation for YFelO M02X ( X = E, H ,S ,C, N, 0, F, where E is
an empty atomic sphere) in the frame of the local spin density (LSD)
functional theory (Yang et ai., 1997b, 1998d). The spin polarized partial
density of the states (DOS) for YFelOMo2andYFelOMo2X is shown in Fig.3.9,
and the calculated magnetic moments are shown in Fig. 3. 10. The atomic
positions of YFelOMo2 and YFelOMo2X (X = E, C, S, N, H, 0, F) are scaled
104 Jinbo Yang and Yingchang Yang

30
20
Fe(8j) Fe(8j)
10
Of---"- of--~~~:s~
-10
-10 E
-20
-30 -20
& 20 20
c::
'0, 10 Fe(8f) 10 Fe(8f)
.!!?
~ Of------L... 0f---4
en
~ -10 -10
Gi'
~ -20 -20
20 20
10 Fe(8i) 10 Fe(8i)
Of----I.. oI----<!:
-10 -10
-20 -20

-1.0 -<l.5 0 0.5 1.0 -1.0 -<l.5 0 0.5 1.0


. Energy (Ry) Energy (Ry)
(a)

20
10 Fe(8j) 10 Fe(8j)

0 o f===l1~a~~;;.j
-10 -10

-20 -20
& 20 20
c:: 10
'0, Fe(8f) 10 Fe(8f)
~ 0 Of---<"
~
~ -10 -10
~ -20 -20
20 20
10 Fe(8i) 10 Fe(8i)
0 Ol----c

-10 -10
-20 I -20 I
I I
-1.0 -<l.5 0 0.5 1.0 -1.0 -<l.5 0 0.5 1.0
Energy (Ry) Energy (Ry)
(b)

Figure 3.9 The electronic structures of YFeloMoz and YFelOMozX ( X=Z. H. B.C.N.O.F).
Continued on page 105.
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)12 Compounds 105

30 I YFe lO Mo2N
20 I
10 Fe(8j) I Fe(8j)
10
o Hr-~~i#l~~~
-10 -10
-20
-20 -30
~ 20 C(2b) 20
c
'5. 10 Fe(8f) 10 Fe(8f)
~
~ 0 t----<:; Ot---~
ci5 -10
~ -10
G5'
~ -20 -20
20 20
10 Fe(8i) 10 Fe(8i)
Or--~ Ot--~

-10 -10
-20 -20 I
I
-30 '---:-'-;:-::-'-:---'!:---::'-:---:-':::--- - 30 '-----:-'-:----::-'7--'c':------::'-:-~---J
-\.0 -D.5 0 0.5 1.0 -1.0 -<l.5 0 0.5 1.0
Energy (Ry) Energy (Ry)
(e)
30 : YFe lO Mo2 0 30 I YFe lO Mo2 F
20 20
I Fe(8j) I Fe(8j)
10 \0 I
o I---l~~~~~~ o f---*--4:-i~~~~
-10 -10
-20 -20
>; -30 -30
0:: 20 20
c
'5. \0 Fe(8f) 10
.!!2 Fe(8f)
~ Or--~ Or----<;..
~ -10 -\0
G5'
~ -20 -20
20 20
10 Fe(8i) \0 Fe(8i)
01---<". Or----<;..
-10 -10
-20 -20
- 30 '---;-'-;:-;;-'';:-~-::-~;-';;---' - 30 '---;-'-;:-;;-'';:-7----;;-~;-';;---'
-1.0 -D.5 0 0.5 1.0 -1.0 -<l.5 0 0.5 1.0
Energy (Ry) Energy (Ry)
(d)

Figure 3.9 Continued from page 104.


106 Jinbo Yang and Yingchang Yang

2.6

-: Fe(8j)
2.4 e: Fe (8f)
'7a A: Fe (8i)
3 .... : Average

".,
E 2.2
0
E
0
.~

5iJ 2.0
~'"

1.8

YFe lO Mo 2 E H B C N 0 F

Figure 3. 10 The calculated magnetic moments of YFelO M02 and YFelO M02 X
(X= E, H, B,C,N,O,F).

according to our experimental results listed in Table 3.4. In order to separate


the geometrical effect of doping YFelO M02 from the chemical bonding effect,
we have performed the following calculations:
( 1) YFelO Moz X (X =
E, C, B, N, H, 0, F) with the coordinates of the
experimental lattice constants of nitrided YFelO MozN.
(2) YFelo Moz with the true experimental values.
(3) YFelO Moz with the experimental lattice constants of the nitrides with
an empty sphere (E) at the 2b interstitial sites. This result represents the
effect of the volume expansion produced by the X atoms. The difference
between (1) and (3) gives the chemical bonding effect of the X atoms.
The band structures of the interstitial modified 1 : 12 compounds and their
host compounds reveal the following tendencies:
( 1) For the nitrides, the magneto-volume effect of the nitrogen atoms
absolutely increases the magnetic moment of each Fe site and the total
magnetic moment of the host compound, and the chemical effect results in an
enhanced moment on the 8i and 8f sites and a reduced moment on the 8j sites,
but also increases the total magnetic moment. The magneto-volume effect
increases the absolute value of the hyperfine fields at three Fe sites, and the
chemical bonding effect will decrease the absolute values of the Fe hyperfine
fields in the 8j and 8i sites, but increase those of the 8f sites. It is found that
the Fe magnetic moments and the hyperfine field at 8i sites are more stable
against the volume expansion than those at the 8j and 8f sites. Both the
chemical bonding effect and magneto-olume effect increase the isomer shift at
the different Fe sites.
=
(2) The interstitial atoms X (X E, H, B, C, N, 0, F, where E is an empty
atomic sphere) atoms influence the electronic structures at the Fe sites
Magnetic Properties and Interstitial Atom Effects in the R( Fe, M) 12 Compounds 107

differently. It is found that the chemical bonding effect of the X atoms is


different, and changes irregularly with the increase of the atomic number of X.
The iron magnetic moments of the Fe (8D sites are reduced by hybridization
with the 2s and 2p electrons of the X atoms (expect X = F). The effect is
greatest for X = B, and decreases toward the end of the series. It was
demonstrated that for X= B, the p states of the B atoms occur energetically in
the lower part of the Fe 3d states, which gives rise to a strong hybridization
effect and a decrease in the magnetic moments. When going from X = B to X =
C, N, 0, F, the potential of the doping atom becomes steeper and steeper,
the p and s states occur at successively lower energies, and thus the
hybridization effect becomes weaker and weaker. Upon going from C to N, 0,
and F, one arrives at a situation which is similar to the case of YFelO M02 E( E
is an empty sphere). As a result, for YFelO M02 F, there is nearly no chemical
bonding effect and hence just an increase of the magnetic moments due to the
magneto-olume effect. It is therefore suggested that F is probably the optimum
doping atom to achieve large magnetic moments in the 1 : 12 compounds,
provided that the sample can be prepared and remain stable.
The increase of the Curie temperature upon nitrogenation can be
understood using band structure calculations. According to the spin fluctuation
theory of Mohn and Wohlfarth (1987),

T SF = M~/(10kBXc), } (3.8)
X~' = (4j.l~)-'[(1/N+ (E F ) + l/N_ (E F ) -2/)J

where Me is the magnetic moment per Fe atom at 0 K, Xc is the enhanced


susceptibility, I is the Stoner parameter, N + (E F ) and N - (E F ) are the spin-
up and spin-down DOS at the Fermi energy, respectively, and T SF is the spin
fluctuation Curie temperature given by
(T c /T sT )2 + Tc/T sF - 1= 0 (3.9)

Here T ST is the Stoner Curie temperature. The parameters as obtained from


the band calculations are listed in Table 3. 9. Generally, T SF indicates the change

Table 3. 9 The band structure parameters of the R (Fe, M) 12 and their interstitial modified
compounds (N + (EF) and N - (EF) are the spin up and down DOS at Fermi level, / is the Stoner
parameter, M o is the magnetization at 0 K, and TSF is the spin fluctuation Curie temperature) .

N+ (E F) N- (E F) Mo TSF(N)/T sF Tc(N)/Tc
/(eV)
(eV- 1Fe- 1) (eV- 1Fe- 1) (Ila/ Fe) (Theor. ) (Exp. )

YFelO Mo2 0.530 0.472 0.96 2.02


YFelOM02H 0.490 0.420 0.97 2.09 1. 10 1. 07
YFelOMo2N 0.423 0.463 0.95 2.20 1. 42 1. 25
NdFe'1 Ti(1) 0.553 0.706 0.90 2.12
(1)
NdFell TiN o. s 0.500 0.646 0.88 2.17 1. 31 1. 32
(1) Jaswal, 1995.
108 Jinbo Yang and Yingchang Yang

of T c' suggesting that an increase of M o and a substantial decrease of N +


(E F ) upon nitrogenation are responsible for the increase of Tc (Sakuma,
1992; Jaswal 1993, 1995).

3.4 Novel Permanent Magnets Based on 1 12 Nitrides

3.4.1 The Domain Structures in Compounds, Nitrides and Carbides


It is well known that the hard magnetic properties are sensitive to the
demagnetization process that is related to the magnetic domain structures. In
order to shed light on the making of permanent magnets with nitrides, it is
important to investigate the domain structures of these materials. The domain
structure investigation was carried out using magnetic force microscopy
(MFM) and Kerr microscopy. The compounds were cut into plates with a
thickness of 2 mm, ground and polished using standard metallurgical
procedures. A suitable grain with a surface nearly normal to its easy axis
[001] was selected for examination using a digital instrument dimension 3000
scanning probe microscopy. Figure 3. 11a shows the atomic force microscopy
(AFM) image of surface for the NdFelO.5 M0 15 compound. The topography of
the sample shows a large number of scratches arising from the polishing
process (dark lines). In addition, there is a small number of elevated regions
(white, circular) which are probably foreign matter left behind after polishing.
A typical domain structure was observed for the NdFelo.5 M0 15 as shown in
Fig. 3. 11 b. This domain structure has two distinct features: extreme
corrugation and introduction of reverse "spike" domains visible as small spots.
This typical domain structure is formed due to a compromise between the
magnetostatic energy and the domain wall energy. Since the total energy per
unit area of the domains is a sum of the magnetostatic energy and the domain
wall energy y, the equilibrium domain structures are obtained by minimizing
this total energy. After observation, this sample was nitrogenated to give a
fully nitrogenated NdFelO.5Mo1.5 sample. Figure 3. 12 shows the MFM image of
NdFelO.5Mo15Nx nitrides. An appearance of stripe domains is observed on the
grain surface of NdFelO.5 Mo1.5 Nx . Obviously, upon nitrogenation, a domain
structure transition is found. This transition is due to a strongly uniaxial
magnetocrystalline anisotropy induced by interstitial nitrogen atoms, because
the wall energy y depends on the magnetocrystalline anisotropy constant K,
being extremely large for these nitrides. In order to decrease the total energy,
a stripe domain is more favorable after nitrogenation. According to the method
of Bodenberger and Hubert ( 1977 ), the average domain width W can be
determined based on the experimental results. The value of W is 0.6 IJm and
Magnetic Properties and Interstitial Atom Effects in the R( Fe, M) 12 Compounds 109

2. 4 IJm for NdFe ,0.5 Mo1.5 and NdFe,D.5 Mo1.5 Nx ' respectively. The average
domain wall energy y and the critical diameter of the single domain particles
Dc' are respectively given by

(3.10)

where f3 has been determined experimentally to be O. 31 (Boenbedrger and


Hubert, 1977) and M s is saturation magnetization. Together with Table 3. 1,
we have estimated and listed the values of y and Dc in Table 3. 10.

o 20.0 J.lrn 0
(a) (b)

Figure 3. 11 The images of a NdFe,o.s Mo1.5. scan taken on a face perpendicular to easy
axis, scan height 40 nm. (a) AFM image showing the topography, (b) MFM image of the
same region.

Figure 3.12 The MFM image of NdFe,o.sMo1.5N.


110 Jinbo Yang and Yingchang Yang

Table 3. 10 Some fundamental magnetic properties of NdFelOs Mo 15 , NdFelOs M0 15 Nx and


NdFelO sMo 1 sC x .
critical diameter
Average Domain-wall Domain-wall Exchange-
of the single
domain width energy thickness constant
domain particles
W y 6 A Dc
2 2 (10-11J/m)
CiJm) (10- J/m ) CA) CiJm)
NdFelO.s M0 15 0.60 1.0 267 2.2 0.22
NdFelO.S M0 15 Nx 2.40 5.7 70 3.2 0.86
NdFelO.s M015 C x 1. 30 3.7 47 2.9 0.97
SmCos (1) 3.00 8.5 51 3.5 1. 60
N~Fe14B (2) 0.85 1.7-2.5 30 1.1 0.15-0.30
(1). (2) Livingston et al.. 1972. and 1987.

The exchange constant A, as well as the domain wall thickness 6 can be


obtained according to the standard continuum model of domain walls (Craik
and Tebble, 1965, Craik, 1970),

y =4 x JA x K} (3.11)
6 = IT x JA/K = lTy/4K
The values of A and 6 calculated from measured values of y and K are
presented in Table 3.9. As shown in Table 3. 9, the domain wall energy and
the critical diameter of single domain particles of the nitrides are much higher
than those of the parent compounds. It is evident that the nitrides have a large
anisotropy constant K and a large exchange constant A, corresponding to a
large y and Dc. As for a comparison of the domain parameters with those of
SmCos and Nd2 Fe14 B, the values of y and Dc of NdFel0.S Mo1.5 Nx are lower
than those of SmCos (Livingston and McConnell, 1972 ) and higher than those
of Nd2 Fe'4 B (livingston, 1985). These results are reasonable, since the
values of A and K in NdFelO.s Mo1.5 Nx lie between the values of SmCos and
Nd2 Fe14 B. Similar domain structures have been observed in NdFelO.s Mo1.5 C x
using Kerr microscopy as shown in Fig. 3.13. The calculated magnetic parameters

Figure 3. 13 The domain structure of the NdFelO.S M0 15 C x


Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)12 Compounds 111

are also listed in Table 3.9. It is evident that the carbides exhibit a lower y
and Dc compared to the nitrides. On the other hand, the domain structure of
the hydrides exhibits little change compared to the parent compounds.

3. 4. 2 Hard Magnetic Properties of Nitrides


The conventional preparation technique for rare-earth transition metal
permanent magnets is the sintering process. However, it is not suitable for
nitrides due to the high sintering temperature. So far mechanical alloying and
rapid quenching have been used to produce hard magnetic powders with high
coercivity. Coercive forces up to 477 - 1035 kA/ m have been obtained by
mechanical alloying techniques (Gong and Hajipanayis, 1993; Yang et al. ,
1996; Zhang et ai., 2000). Melt spinning of NdFelO M02Nx results in
coercivities of 477 and 701 kA/m (Tang et ai., 1992, 1993). A coercivity of
438 kA/m for nitriding ribbons has been reported (Pinkerton et ai., 1994). A
coercivity of 828 kA/m for NdFelO.5 Mo1.5 Nx has been achieved by optimizing
the annealing and nitriding conditions (Yang et ai., 2000b). The
hydrogenation, disproportion, desorption and recombination (HDDR) process
was also used to prepare R(Fe,M),2Nx(Yang et ai., 1994b; Sugimoto et ai.,
1994; Jurczyk et ai., 1995; Jin et ai., 1998; Oleinek et ai., 1999b).
However, the powders prepared using mechanical alloying, melt spinning and
HDDR are isotropic, thus the remanence and the maximum energy product are
low. Therefore, it is necessary to develop a new procedure for preparing the
1 : 12 nitrides concerning the formation conditions of the high remanence and
the creation of a high coercivity.
In this part, RFelO.5 M1.5 Nx (R = Nd, Pr, M = V, Mo) powders with
favorable magnetic properties are obtained by optimizing the homogenizing
treatment of the parent compounds, the nitrogenation temperature and time,
and the particle size before and after nitrogenation. The key problems dealing
with the manufacturing of maghets and the prospects of for the 1 : 12 nitrides
are discussed.
The typical X-ray diffraction patterns of the NdFelO.5 V1.5 and NdFelO.5 V1.5Nx are
given in Fig. 3. 14. The X-ray patterns identify that the compounds are of the
1 : 12 single phase, which also can be verified from the thermomagnetic
curve. There is a little precipitation of a-Fe after nitrogenation and its content
increases with increasing nitrogen content. All the X-ray diffraction peaks of
NdFelO.5 V1.5 Nx shift to lower angles as x is less than 2.0, suggesting that the
nitrogen enters into interstitial sites and results in the lattice expansion. The
positions of the diffraction peaks are nearly unchanged after x > 2. 0,
indicating the nitrogen does not enter the interstitial sites further, and at the
same time a-Fe and neodymium nitrides are formed. In addition, as compared
to the RFe" Ti and R(Fe, Mo) 12 compounds, the R(Fe, V) 12 alloys possess a
stronger ability to absorb nitrogen atoms. The content x in the RFelO.5 V1.5 Nx
112 Jinbo Yang and Yingchang Yang

a-Fe

30 40 50
2en
Figure 3.14 The X-ray diffraction patterns of NdFelO.5 V1.5 Nx .

ranges from 0 to as large as 2. 7, which is much more than the theoretical


value of one N per formula. One point in understanding this is that the
measurements include everything gained during nitrogenation such as nitrogen
in the neodymium nitride precipitates, nitrogen in the particle surface
(physical absorption) or in the bulk, and the formation of Fe-N compound as
well. In fact, during the nitrogenation process, it is inevitable that a small
amount of impurity phases such as iron and rare-earth nitrides are formed.
Thus the actual N content in the 1 : 12 phase may be significantly less than the
measured value by mass. Because the nitrogen content has so great an
influence on the crystal structure and the phase composition of the nitrides, it
will finally affect the magnetic properties of the magnetic powders. As a
comparison, the magnetization curves of RFelO.5 V1.5Nx with different nitrogen
contents at room temperature, parallel and perpendicular to the alignment
direction, are plotted in Fig. 3. 15. The corresponding intrinsic magnetic
parameters are listed in Table 3.11. From Fig. 3. 5 and Table 3. 11, it is seen
that the magnetic anisotropy and the saturation magnetization change
remarkably with nitrogen content x. The magnetocrystalline anisotropy field
(H.) and saturation magnetization (M s ) increase steadily with increasing
nitrogen content and the former reaches a maximum value at x = 2. O. Noting the
increasing of H., it can be concluded that NdFelO.5 V1.5 is not fully nitrogenated until
x reaches 2 or so. With further nitrogenation, the rnagnetocrystalline anisotropy field
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)'2 Compounds 113

1.2

0.5 1.0 1.5 2.0


J10H (T)
Figure 3. 15 The magnetization curves of NdFelO.5 V1.5Nx parallel (open symbols) and
perpendicular (closed symbols) to the alignment direction at room temperature.

Table 3.11 The intrinsic magnetic properties, saturation magnetization (J./o M s ), anisotropy
constants (K I' K l ) and anisotropy field (J./o H.) at room temperature for NdFelO 5VI 5N x .
J./oMs KI K2 J./o H.
x
(T) 3 3
(MJ/m ) (MJ/m ) (T)
0.0 1.146 0.13 0.05 0.5
0.7 1. 212 3.34 0.58 7.3
1.4 1.266 3.52 0.62 9.9
2.0 1.321 3.80 0.71 10.4
2.7 1.327 -0.11 3.33 -

decreases drastically due to the poor crystal structure. This is consistent with
the X-ray pattern of x = 2. 7, where a certain amount of impurity phases are
formed (Oleinek et al. , 1998). Thus, a proper nitrogen content with a small
amount of impurity phases results in powders having the desired hard magnetic
properties. In the case of insufficient nitrogenation, the core of the large grains
remains unnitrided, even though the small particles are fully nitrided (Yang et
aI., 1998b), which gives rise to lower coercivity and magnetization. This is
consistent with the 2 : 17 compounds (Skomski et aI., 1993a, 1993b; Zhang et
aI., 1997). Figure 3. 16 shows the typical domain structures of the NdFelO.5 Mo1.5
particles with differing nitrogenation time. It is evident that a soft-core region
is formed in the partly nitrided particles. These results suggest that both
insufficient nitrogenation and over-nitrogenation are harmful for the intrinsic
magnetic properties. Figure 3. 17 shows the variation of the intrinsic coercivity
I Heand the remanence Mras a function of x for NdFelo.5V1.5N., where x is the

nominal nitrogen content. A maximum in I He and M r is observed around x =


2. 0, which confirmed that the nitrogen content is essential for high performance
powders.
114 Jinbo Yang and Yingchang Yang

(b) 4h

Figure 3. 16 The domain structures of the NdFelo.5 Mo1.5 with different nitrogenation times.

Generally, the nitrogen content is dependent on the nitrogenation


process, so the nitrogenation temperature and time playa key role in obtaining
good hard magnetic properties. By optimizing the nitrogenation temperature
and time, the highest coercivity of NdFel0.5 V1.5Nx is obtained with
nitrogenation at T= 510 C for 8 h. The hysteresis loop of these powders is
Magnetic Properties and Interstitial Atom Effects in the R( Fe, M) 12 Compounds 115

presented in Fig. 3. 18. The room temperature permanent magnetic properties


of NdFelO.5 V1.5Nx are: CBH)max = 137 kJ/m 3 C17.2 MGOe) , B, = 1.08 T, and
,He = 335 kA/m C4. 3 kOe). The intrinsic coercivity can be enhanced to
340. 5 kA/ m at the cost of remanence and energy product Ccorresponding
values are B, = 1. 05 T, CBH)max = 128 kJ/m3 ). In addition, ,He of the
powders exhibits a large dependence on the magnetization field. For example,
coercivities of 318 kA/m and 398 kA/m are obtained under applied fields of 2
and 4 T for NdFelO.5 V1.5N x . Coercivities of 406 and 502 kA/m are obtained
under fields of 2 and 4 T, respectively, for NdFelO.5 Mo1.5 Nx CYang et al. ,
1994).
1.2

1.1

E
1.0 ::;t
:f

0.9

L -_ _---'- -'--_ _---'-_----' 0.8


1.5 2.0 2.5
Nitrogen content, x

Figure 3. 17 The dependence of coercivity and remanence on the


nitrogen content in NdFelO.5 V1.5Nx compounds.

,,
1.5

1.0
....... ,,
1.5 K ..;.
:..~~~
.
...

,,
I RT

~ 0.5 I

-----------~
.
c:
o o
:~v --f--

...
51 -0.5 :
~'" :,
,,
-1.0
.......'""l::
-1"-
I
-1.5 L---L.._-...J'--_-'-'_ _-'--_--'-_
-2 -I 0 2
lloH (T)
Figure 3. 18 The hysteresis loops of NdFelO.5 V1.5 Nx at room temperature and 1.5 K.
116 Jinbo Yang and Yingchang Yang

Coercivity and remanence are extrinsic magnetic properties, which are


strongly dependent on the technical process. The dependence of the coercivity
and remanence on the particle size of the NdFelO.s V1.5N. powders is illustrated
in Fig. 3.19. The coercivity first increases as the particle size decreases, and
reaches a maximum value at a size of about 1 \.1m. As the particle size
decreases further, the coercivity begins to decrease. The initial increase of
coercivity can be explained by the single domain theory and the nucleation
model. The remanence has the same trend with respect to the particle size,
but reaches a maximum value at a size of about 3 \.1m. The transmission
electron diffraction pattern of powders with a maximum coercive force
indicates that the fine powders are single crystals. With prolonged milling, the
powders are found to be defective and are transferred into polycrystals (Yang
et aI., 1998e). The results indicate that the crystal structure of the powders
was damaged after prolonged grinding, which drastically degrades I He and M,
of the nitrides. A similar case has also been found in NdFelo.s Mo1.5 N. The
temperature dependence of I He determined by measuring the temperature
dependence of the hysteresis loops in a temperature range from room
temperature to 500 K for NdFelo.sV1.5N. and NdFelo.sMo1.5N. powders is shown
in Fig. 3.20. The temperature coefficient of coercivity is smaller than that of
the Nd2Fe148 magnets and close to that of Sm2Fe17N. magnets (Suzuki et al. ,
1993).
350 1.2
~

300 I \
\
1.1

\ 1.0
I ~ 110M,
:g 250 r "-
"- ....... 0.9 E
~ I
~ 200
I
r
0.8 1
I 0.7
150 I
I 0.6
~
100 0.5
2 3 4 5
Size (11m)

Figure 3. 19 The dependence of room temperature coercivity and


remanence on the particle size of NdFel0.5 V1.5 N.

Similar efforts have been made to produce anisotropic magnetic powders of


NdFelO.sMo1.5N., PrFel0.sMo1.5N. and Pr(Fe, V)12N. (Yang et aI., 1997a; Mao et
aI., 1997, 1998c, 2001). Figure 3.21 shows the hysteresis loops of some nitride
powders. The magnetic properties are summarized in Table 3. 12. Using these
powders, polymer bonded magnets with a density of 5.5 - 6.0 g/cnT and a
Magnetic Properties and Interstitial Atom Effects in the R( Fe, M) 12 Compounds 117

800

.: NdFelO.5M0I.5Nx
600
.: NdFe lO.sVI.5Nx
~

C 400
~

200

300 350 400 450


T(K)
Figure 3. 20 The temperature dependence of the coercivity of the
NdFe,o.5 Y1.5 N. and NdFe,o.5 Mo1.5 N. magnets.

maximum energy product (BH)rrex =


72 -79.6 kJ/rrr (9-10 MGQe) have been
prodJced in large scale. As a matter of fact, R(Fe, V)'2 N. (R = Nd, Pr) nitrides
exhibit a larger saturation magnetization and a much higher Curie temperature than
those of R(Fe, Mo)'2N. PrFelO.7S V1.2SNx show a saturation magnetization of Po M. =
1. 54 T, and the Curie temperature Tc =
800 K, and an estimated theoretical
maximum energy prodJct of 472 kJ/ rrr (59 MGOe) (Mao et al., 2001). The
excellent intrinsic magnetic properties, together with a lower ratio of the rare-
earth to the transition metal and the lower price of Pr, renders the Pr(Fe,V)12N. a
favorable candidate system for applications in the inexpensive bonded magnets
and a promising competitor to the Nd2Fe'4B magnets.

Table 3. 12 Intrinsic coercivity ,He' Remanence B r , maximum energy product (BH)max at


room temperature and 1.5 K, and the Curie temperature Tefor R(Fe,M)'2N. powders.
,He(kA/m) Br(T) (BH)max (kJ/m 3 ) Tc(K)

at room temperature
NdFe,o.5 Mo,. 5N. 478 1. 02 169 665
PrFe,0.5 Mo1.5 Nx 334 0.98 132 650
NdFe,o.5 y 15 N. 335 1. 08 137 >860
PrFe'O.75 Y1.25N. 232 1. 18 135 >820
PrFe'O.5 Y1.5 N. 259 1. 09 129 >820
1.5K
NdFe,o.5 Mo1.5 N. 3049 1. 24 255
NdFe,o.5 Y1.5 N. 1791 1. 22 275
PrFe'O.5 Y1.5N. 1592 1. 23 229
118 Jinbo Yang and Yingchang Yang

1.5 NdFe lO .5 M0 I.5 Nx

I.5K RT
E
1c 0.5

.~ o ------------ ----r--- -------------.


.~., I
I
I
I
~ -0.5 I
I

:z:'" I
I
I
I
-1.0 I

-1.5 L - . . . L - _ - - - l L - . _ - ' -_ _- ' - - _ - - ' _


-2 2

1.5

1.0
I RT
I

1
I
I
I
0.5 I
I
co I
I
o --- ---------- --r- ----------- --_.
:~., I
I
I
I
~ -0.5 I
I

:z:'" I
I
I
I

-1.5 '-----'--_ _'---_-'-_ _-'-_----'_


-2 2

Figure 3. 21 The hysteresis loops at room temperature and 1. 5 K for


(a) NdFelO.sfv101.5Nx and (b)PrFelO.SV1.5Nx.

3. 5 Conclusions and Prospect

These results show that by understanding the intrinsic and external magnetic
properties of interstitial modified 1 : 12 compounds, hard magnetic materials
with interesting magnetic properties can be prepared. In fact, a polymer
bonded magnet shows a maximum energy product of 72 - 79. 6 kJ/ m3 at room
temperature has been produced recently, which is a good prospect in the
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)'2 Compounds 119

technological applications. Compared to commercial Nd2Fe14 B powders, they


have a moderate coercive force, higher Curie temperature, higher stability,
larger maximum energy product, and are suitable for some practical
applications. 1 : 12 nitrides also have the advantage that the hard magnetic
properties at low temperatures are much better than those at room
temperature, in as much as the easy magnetization direction of Nd2 Fe14 B
changes to the easy-cone at low temperature. Further decreasing the third
element M content in the host compounds will increase the saturation
magnetization and the Curie temperature, resulting in a higher energy product
(BH)max' A suitable quaternary substitution might improve the coercivity and
the saturation magnetization. It would also be worthwhile to prepare
nanocomposite R (Fe, M)12 Nx/cx-Fe (Fe-Co) exchange-coupling magnets,
which might be the alternative approach to yield a further improvement in the
permanent magnetic properties of the 1 : 12 nitrides.

References
Anderson,O.K. Phys. Rev. B 12: 3060 (1975)
Anagnostou, Z. , M. C. Christides and N. Niarchos. Solid State Commun. 78:
681C 1991)
Asano, S. , S. Ishida and S. Fujii. Physica B 190: 155 (1993)
Bodenberger, R. and A. Hubert. Phys. Status Solidi A 44: K7(1977)
Buschow, K.H.J. Rep. Prog. Phys. 40:1179(1977)
Buschow, K.H.J. J. Appl. Phys. 63: 3130(1988)
Buschow, K.H.J. J. Magn. Magn. Mater. 100: 79(1991)
Coey, J. M. D., H. S. Li, J. P. Gavign, J. M. Cadogan and B. P. Hu. In:
Mitchell I. V. , J. M. D. Coey, G. Givord, I. R. Harris and R. Hanitsh eds. ,
Concerned European Action on Magnets. Elsevier London: p. 76 ( 1989)
Coey, J.M.D. and H. Sun. J. Magn. Magn. Mater. 87: L251(1990)
Craik, D. J. and R. S. Tebble. Ferromagnetism and Ferromagnetic
Domains. Wiley, New York, (1965)
Craik, D.J. Contemp: Phys. 11: 65(1970)
Croat, J.J., J.F. Herbst, R.W. LeeandF.E. Pinkeron. J. Appl. Phys. 55:
2078(1984)
Endoh, M., K. Nakamura and H. Mikami. IEEE Trans. Magn. 28: 2560
(1992)
Fernando, A. S. , J. P. Wood, S. S. Jaswal, B. M. Patterson, D. Welipitiya,
A.S. Nozareth and D.J. Sellmyer. J. Appl. Phys. 73: 6319(1993)
Fruchart, D., O. Isnard, S. Miraglia, L. Pontonnier, J. L. Soubeyroux and
R. Fruchart. J. Alloys Compo 203: 157(1994)
Fujii, H. and H. Sun. In: Buschow K. H. J. , ed. Handbook on Magnetic
Materials vol. 9 ed. Elsevier, pp. 304 - 502; Amsterdam ( 1995) and
references there in
Gong, W. and G. C. Hadjipanayis. IEEE Trans. Magn. 28: 2563(1992)
120 Jinbo Yang and Yingchang Yang

Gong, W. andG.C. Hadjipanayis. J. Appl. Phys. 73: 6245(1993)


Goto, T., Y. Takahashi and S. Abe. J. Magn. Magn. Mater. 140-144:
1021(1995)
Hadjipanayis, G.C., R.C.Hazelton and K.R., Lawless. J. Appl. Phys. 55:
2073(1984)
Hu, B. P. , E. D. Wendhausen, W. Pirshke, A. Handstein and K. - H. MUlier.
IEEE Trans. Magn. 30: 645(1994)
Hu, W. Y. , J. Z. Zhang, Q. Q. Zheng and C. Y. Pan. J. Appl. Phys. 76:
6751(1994)
Hu, Z. , W. B. Yelon, X. Zhang and W. J. James. J. Appl. Phys. 79: 5522
(1996)
Hurley, D. P. F. and J. M. D. Coey. J. Phys.: Condens. Matter 4: 5573
(1992)
Iriyama, T., K. Kobayashi, N. Imaoka, T. Fukuda, H. Kato and Y.
Nakagawa. IEEE Trans. 28: 2326(1992)
Ishida, S., S. Asano and S. Fujii Physica B 193: 66(1994)
Izumi, H., Y. Seyama, K. Machida and G. Adachi. J. Alloys Compo 233:
231( 1996)
Jaswal, S. S., W. B. Yelon, G. C. Hadjipanayis, Y. Z. Wang and D. J.
Sellmyer. Phys. Rev. Lett. 67: 644 (1991)
Jaswal, S. S. Phys. Rev. B 48: 6156 (1993)
Jaswal, S. S. In: F. Granjean , G. J. Long, and K. H. J., Buschow, eds.
Interstitiallntermetallic Alloys. Kluwer Academics Publisher, Netherlands,
pp. 411- 432( 1995)
Jin, Z. Q., W. Tang, J. R. Zhang, L. Lu, S. L. Tang and Y. W. Du. J.
Magn. Magn. Mater. 187: 231 (1998)
Jurczyk, M., P. B. Gwan, S. J. Collocott. J. Alloys Compd. 221: 114
(1995)
Li, Z.W., X.Z. ZhouandA.H. Morrish. J. Phys.: Condens. Matter 4: 10,
409(1992)
Li, Z.W., X.Z. ZhouandA.H. Morrish. J. Phys.: Condens. MatterS: 3027
(1993)
Li, H. S. and J. M. D. Coey,ln: J Buschow, K. H. , ed. Handbook of Magnetic
Materials. Vol. 6. Elsevier Science Published B. V. (1991)
Li, H.S. and J.M. Cadogan. J. Magn. Magn. Mater. 109: L153(1992)
Li, Y.P. and J.M.D. Coey. Solid State Commun. 81: 447 (1992)
Livingston, J. D. and M. D. McConnell. J. Appl. Phys. 43: 4756 (1972)
Livingston, J. D. J. Appl. Phys. 57: 4137 (1985)
Mao, W. H. , B. P. Cheng, J. B. Yang, X. D. Pei and Y. C. Yang. Appl.
Phys. Lett. 70: 3044(1997)
Mao, W. H. , J. B. Yang, B. P. Cheng and Y. C. Yang. J. Appl. Phys. 83:
6640(1998a)
Mao, W. H. , J. B. Yang, B. Cui, B. P. Cheng, Y. C." Yang et al. J. Phys.:
Condens. Matter 10: 2611(1998b)
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)12 Compounds 121

Mao, W. H. , J. B. Yang, B. P. Cheng, Y. C. Yang, H. L. Du, B. S. Zhang,


C.T. Ye, J.L. YangandS. L.Ge. Phys. Lett. A243: 163(1998c)
Mao, W. H. X. Zhang, C. Ji, H. Chang, B. P. Cheng. Y. C. Yang, H. Du,
Y. Xue, B. Zhang, L. Wang and F. Li. Acta Mater. 49: 721(2001)
Mishra, R. K., G. Thomsa. T. Yoneyama, A. Fukuno, and T. Ojima. J.
Appl. Phys. 53: 2517 ( 1981) .
Mohn, P. and E. P. Wohlforth. J. Phys. F 16: 2421(1987)
MOiler. K.-H., D. Eckert, F. Fischer, A. Handstein, D. Hinz, R. Kratz, H.
Krug, P. Oleinek, S. Siegel, P. Verges, M. Wolf and J. B. Yang. Physica
B 294- 295: 203 (200 1)
Obbade, S. , S. Miraglia. , D. Fruchart and P. L'Heritier. Z. Phys. Chem.
163: 161(1988)
Ohashi, K. , Y. Tawara. R. Osugi and M. Shimao. J. Appl. Phys. 64: 5714
(1988)
Ojima T. , S. Tomizawa, T. Yoneyama and T. Hori. Jpn J. Appl. Phys. 16:
691(1977)
Oleinek, Ph., W. Kockelmann, K.-H MOiler, M. Loewenhaupt and L.
Schultz. J. Alloys Compd. 281:306(1998)
Oleinek, Ph, D. Eckert, K-H. MOiler and L. Schultz. J. Phys. D: Appl.
Phys. 32: 1578(1999a)
Oleinek, Ph., O. Gutfleisch, A. Handstein, K.-H. Muller. L. Schultz and I.
R. Harris. IEEE Trans. Magn. 35: 3247 ( 1999b)
Pinkerton, F. E., C. D. Fuerst and J. F. Herbst. J. Appl. Phys. 75: 6015
(1994)
Oi, O. N. , Y. P. Li and J. M. D. Coey. J. Phys.: Condens. Matter 4: 8209
(1992)
Oi, O.N., B.P. Hu and J.M.D. Coey. J. Appl. Phys. 75: 6235(1994)
Sagawa. M. , S. Fuj imura, M. Togawa and Y. Matuura. J. Appl. Phys. 55:
2083(1984)
Sakuma, A. J. Phys. Soc. Jpn. 61: 4119 (1992)
Schultz, L. and M. Katter. In: Long. G. J. and F. Granjean, eds.
Supermagnets and Hard Magnetic Materials. Kluwer Academics Publisher.
Netherland, p. 227 ( 1991 a)
Schultz. L. , K. Schnitzke, J. Wecker, M. Katter and K. Kuhrt. J. Appl.
Phys. 70: 6339(1991b)
Sellmyer, D. J., A. Ahmed, G. Muench and G. C. Hadjipanayis. J. Appl.
Phys. 55: 2088(1984)
Sinnemann, T., K. Erdmann, M. Rosenberg, K. H. J. Buschow. Hyperfine
Interact. 50: 675( 1989a)
Sinnemann, T., M. Rosenberg and K. H. J. Buschow. J. Less-Common
Met. 146: 223(1989b)
Skolozdra, R. V., E. Tomey, D. Gignoux, D. Fruchart and J.L. Soubeyroux.
J. Magn. Magn. Mater. 139: 65(1995)
Skomski, R., K. H. MOiler, P. A. Wendhausen and J. M. D. Coey. J. Appl.
122 Jinbo Yang and Yingchang Yang

Phys. 73: 6047 (1993a)


Skomski, R. and J.M.D. Coey. J. Appl. Phys. 73: 7602 (1993b)
Strnat K. J., G. Hoffer, J. Olson, W. Ostertag and J. Becker. J. Appl.
Phys. 38: 1001(1967)
Sugimoto, S. , T. Tatsuki, H. Nakamura, M. Okada and M. Homma. Trans.
Mater. Res. Soc. Jpn. 14B: 1041(1994)
Sun, H., M. Akayma, K. Tatami and H. Fujii. Physica B 183: 33(1993a)
Sun, H., Y. Morij, H. Fujii, M. Akayama and S. Funahashi. Phys. Rev.
B 48: 13,333(1993b)
Sun, J.J., D.S. Xue, B.G. Shen and F.S. Li. Phys. Status Solidi A 136:
K55(1993c)
Suzuki, S. , T. Miura and M. Kawasaki. IEEE Trans. Magn. 29: 2815(1993)
Suski, W. In: Gscheidner, K. A. Jr. and Eyrung L. Handbook on the Physics
and Chemistry of Rare Earth, Vol. 22. Elsevier Science Published B. V.
(1996)
Tang, Z. X. , E. W. Singleton and G. C. Hadjipanayis. IEEE Trans. Magn. 28:
2572( 1992)
Tang, Z. X., G. C. Hadjipanayis and V. Papaefthymiou. J. Alloys Compo
194: 87 (1993)
Tomey, E. , M. Bacmann, D. Fruchart, D. Gignoux and J. L. Soubeyroux. J.
Magn. Magn. Mater. 157/158: 113(1996)
Wang, Y.Z. and G.C. Hadjipanayis. J. Appl. Phys. 70: 6009(1991)
Wang, Y. Z. , B. P Hu, X. L. Rao, G. C. Liu, L Yiu, W. Y. Lai , W. Gong
and G.C. Hadjipanayis. J. Appl. Phys. 73: 6251C1993a)
Wang, Y. Z., G. C. Hadjipanayis, Z. X. Tang, W. B. Yelon, V.
Papaefthymiou, A. Moukarika and D. J. Sellmyer. J. Magn. Magn. Mater.
119: 41C1993b)
Wei, Y. N. , K. Sun, Y. B. Fen, J. X. Zhang, B. P. Hu" Y. Wang, X. L.
Rao and G. C. Liu. J. Alloys Compo 194: 9(1993)
Yang Y. C. , B. Kebe, W. J. James, J. Deportes and W. B. Yelon. J. Appl.
Phys. 52: 2077(1981)
Yang, J., S. Dong, Y.C. Yang and B. P. Cheng. J. Appl. Phys. 75: 3013
( 1994a)
Yang, J., S. Dong, W. Mao, P. Xuan and Y. C. Yang. J. Appl. Phys. 76:
6053(1994b)
Yang, J. , Ou. Mao and Z. Altounian. J. Appl. Phys. 79: 5519(1996)
Yang, J.B., W.H. Mao, B.P. Chen, Y.C. Yang, H. Xu, H. B. Han, S.L.
Ge and W. J. Ku. Appl. Phys. Lett. 71: 3290(1997a)
Yang, J. B. , W. H. Mao, Y. C. Yang, S. L. Ge, D. F. Cheng. Phys. Rev.
B 56: 15,647(1997b)
Yang, J.B., B. Cui, W.H. Mao, B.P. Cheng, J. Yang, B. Hu, Y C. Yang
and S. L. Ge. J. Appl. Phys. 83: 2700 (1998a) .
Yang, J.B., W.H. Mao, Y.C. Yang, S.L. Ge, Z.J. Zhao and F.S. Li. J.
Appl. Phys. 83: 6923(1998b)
Magnetic Properties and Interstitial Atom Effects in the R(Fe,M)12 Compounds 123

Yang, J. B. , W. H. Mao, Y. C. Yang, S. L. Ge, Z. J. Zhao and F. S. Li. J.


Appl. Phys. 83: 1983(1998c)
Yang, J. B. , W. H. Mao, Y. C. Yang and S. L. Ge. J. Phys. Soc. Jpn. 67:
576(1998d)
Yang, J.B., B. Cui, W.H. Mao, B.P. Cheng, Y.C. Yang and S.L. Ge. J.
Magn. Magn. Mater. 182: 131C 1998e)
Yang, J.B., Ph. Oleinek, D. Eckert, M. Wolf and K.-H Muller. J. Magn.
Magn. Mater. 214: 44 (2000a)
Yang, J.B., Ph. Oleinek and K.-H Muller. J. Appl. Phys. 88: 988 (2000b)
Yang Y. C., X. D. Zhang, S. L. Ge, L. S. Kong and Q. Pan. In: S. G.
Sankar ,ed. proceeding of the sixth Symposium on Magnetic Anisotropy and
Coercivity in Rare Earth Metal Alloys. Carnegie Mellon University,
Pittsburg,p. 190( 1990)
Yang, Y.C., X. Zhang, S.L. Ge, Q. Pan, L. Kong, H. Li, J.L. Yang. B.
Zhang, Y. Ding, C. Yeo J. Appl. Phys. 70: 6001C1991a)
Yang, Y.C., X.D. Zhang, L.S. Kong and S.L. Ge. Appl. Phys. Lett. 58:
2042(1991b)
Yang, Y. C., X. Pei, H. Li, X. Zhang, L. Kong, Q. Pan, M. Zhang. J.
Appl. Phys. 70: 6574(1991c)
Yang, Y.C., X.D. Zhang, L.S. Kong, Pan Q., J.L. Yang, Y.F. Ding, B.
S. Zhang and C. T. Yeo Solid State Commun. 78: 313(1991d)
Yang, Y. C. , Q. Pan, X. Zhang, J. Yang, M. Zhang and S. L. Ge. Appl.
Phys. Lett. 61: 2723(1992)
Yang, Y. C. , Q. Pan, X. D. Zhang, M. H. Zhang, C. Yang, Y. Li, S. L. Ge
and B. Zhang. J. Appl. Phys. 74: 4066(1993a)
Yang,Y.C., X.D. Zhang, S. Z. DongandQ. Pan. J. Appl. Phys. 74: 6847
(1993b)
Yang, Y. C. , Q. Pan, B. P. Chen, X. D. Zhang, Z. X. Liu, Y. X. Sun and S.
L. Ge. J. Appl. Phys. 76: 6725(1994)
Yelon, W.B. and G.C. Hadjipanayis. IEEE Trans. Magn. 28: 2316(1992)
Zhang, L., S. G. Sankar, W. E. Wallace and S. K. Malik. Proc. Int.
Workshop on Rare-Earth Magnets and Their Applications, 11 th: 2493
(1990)
Zhang, X. D., B. P. Cheng and Y. C. Yang. Appl. Phys. Lett. 77: 4022
(2000)
Zhang, Y. D., J. I Budnick, W. A. Hines and J. M. Gromek. J. Phys.:
Condens. Matter 9: 120(1997)

The authors are grateful to Dr. W. J. James at Graduate Center for Materials Research,
University of Missouri-Rolla, USA for assistance with the manuscripts. This work was
partially supported by State Key Project of Fundamental Research and National Science
Foundation of China.
4 Nanocrystalline Soft Magnetic Materials and
Their Applications

Yoshihito Yoshizawa

4. 1 Introduction

Nanocrystalline soft magnetic materials show attractive magnetic properties


and unique microstructure. Hence, research interest in nanocrystalline soft
magnetic materials has increased. FeCuM(Si)B alloys (M: Nb, Ta, Mo, W,
Zr, etc.) (a trade name "FINEMET") known as a pioneer material in
nanocrystalline soft magnetic materials (Yoshizawa et aI., 1988a, 1988b;
1989a, 1990a, 1991a) consists of bcc Fe(-Si) grains of 5-30 nm in grain
Yoshizawa and Yamauchi, diameter, which are extremely fine compared with
those of conventional crystalline soft magnetic materials. They are prepared
by heat treatment for crystallizing the amorphous FeCuM(Si)B alloys made by
rapid solidification or deposition techniques.
After the first report of the FeCuNbSiB alloys (Yoshizawa et al. , 1988a),
the various nanocrystalline soft magnetic materials were reported by similar
processing (Kataoka et al., 1989; Hasegawa and Saito, 1990a, 1990b,
1992; Taneko et aI., 1991; Nakanishi et aI., 1991; Sawa and Takahashi,
1990; Suzuki et aI., 1990,1991; Fujii et al. , 1991; Watanabe et aI., 1993;
Tomida, 1994).
The Fe-based nanocrystalline soft magnetic materials can be classified
into typical three basic systems:
(1) Fe(Au, Cu)M(Si)B (M: Ti, Zr, Hf, V, Nb, Ta, Mo, W) They
consist of a body centered cubic (bcc) Fe (-Si) phase, a residual amorphous
matrix phase and an face centered cubic (fcc) Cu (-Fe) or an fcc-Au (-Fe)
phase.
(2) FeMB (M: Ti, Zr, Hf, V, Nb, Ta, Mo, W) They consist of a bcc-
Fe phase and a residual amorphous phase.
(3) FeM(C, N) (M: Ti, Zr, Hf, V, Nb, Ta, Mo, W) They are mainly
produced by thin film process such as sputtering. They consist of a bcc-Fe
phase, a dispersed nano-meter size carbide or nitride phase, which exists at
grain boundary triple points.
These carbide phase and nitride phase tend to pin bce-Fe grain boundaries.
The nanocrystalline Fe-based alloys show the high saturation magnetic
Nanocrystalline Soft Magnetic Materials and Their Applications 125

induction of 1. 0 - 1. 7 T and extremely high initial relative permeability of


about from 1 x 10 4 to 2 x 10 5 comparable to Co-based amorphous materials or
supermalloy.
Especially, in the nanocrystalline soft magnetic materials, the FeCuNbSiB
alloys were first commercialized. They have extended the market as the
magnetic materials for the electromagnetic compatibility (EMC), power
electronics applications and so on.
In this chapter, we review the studies on magnetic properties,
microstructure, and applications in the nanocrystalline FeCuMSiB (M: Nb,
Ta, Mo, W, Zr, etc. ) soft magnetic materials.

4. 2 Magnetic Properties and Microstructure

4.2.1 Magnetic Properties


Figure 4. 1 shows the relationship between relative permeability and saturation
magnetic induction for typical soft magnetic materials (Yoshizawa, 1999).
The nanocrystalline soft magnetic materials exhibit high permeability from
several tens of thousands to about 2 x 10 5 combined with the high saturation
magnetic induction of from 1. 0 to 1. 7 T.

Co-based amorphous alloy f= I kHz


Permalloy
Nanocrystalline soft
magnetic materials

::t
A limit for

~
conventional
4 materials
10

',/
Mn-Zn ferrite
,,
3
10
0 0.5 1.0 1.5 2.5
Bs(T)

Figure 4. 1 Relationship between relative permeability IJ, and saturation magnetic


induction B s for typical soft magnetic materials.

Table 4. 1 gives the magnetic properties of the typical Fe-based nanocry-


stalline soft magnetic materials. Thus, the development of the nanocrystalline
126 Yoshihito Yoshizawa

soft magnetic materials made it possible to combine high saturation magnetic


induction and high permeability.

Table 4. 1 Magnetic properties of the typical Fe-based nanocrystalline soft magnetic


materials.
Composition Cat. %) 8 s (1) HcCA/m) iJr x 10 4 at f= 1 kHz

Fens Cu, Nb3Sh3 5B9 1. 24 0.5 10 to 17


Fens CUI Nb3Si ,6.5B6 1.18 1. 1 7.5
Fes, Cu, Nb3Si 2B'3 1. 55 12.8 0.9
FenCUo.6 Nb2.4Sill B9 1.45 0.8 15.7
Fea1.5 Cu, Nb3Si2B'2.5 1. 56 10.4 1.4
Fe7S.3 CUO.6 Nb2.6Si9.5 B9 1. 50 1.0 10.9
Fe79 5CUI Nb3Si 4B ,05 P2 1.5 2.6
Fens Cu, Nb3Ge,3.5 B9 1.0
Fe79.a CUO.6 Nb2.6Sis B9 1. 56 5.2 2.7
Feso Cu, Zr2 Si, B9 1.60 5.0 2
FeS5 CUI Zr5 B9 1. 54 1.7 3.87
Fen.5 Cu, Ta3 Sh3.5 B9 1. 14 1.3 4
Fen5CU, M03Si'3.5 B9 1. 21 1.1 7
Fen 5Cu, W3Sh3.5 B9 7
Fens Cu, Zr3 Si 13 . 5B9 1. 20 1.3 7.5
FeS6 Zr7 B6Cu, 1. 52 3.2 4.1CH m =0.8A/m)
FeS9Zr7 B4 1. 57 13.0 1.23
Fe9' Zr7 B2 1.70 7.2 1.4CH m =0.8A/m)
Fes9Hf7 B4 1. 59 5.3 1.85CH m =0.8A/m)
Fe79Nb7B,4 1. 24 12.2 1. 36C H m = O. 8A/m)
FeS4 Nb7B9 1.49 4.2 1.9CH m =0.8A/m)
FeS6 Zr4 Nb3B6Cu, 1. 53 3.7 1.8CHm =0.8A/m)
Fe90 Hf7B2CU, 1.6 3.0 1.8CH m =0.8A/m)

FeS4 Zr3.5Nb3.5 BaCu, 1. 53 1.7 10CHm =0.8A/m)

The magnetic properties of these materials are controlled by alloy


composition and annealing conditions.
Figure 4.2 shows the Cu content x dependence of relative permeability /J,
of Fe74.5-xCuxNb3Si,3.sBg alloys annealed at 823 K for 3.6 ks. By adding Cu to
the FeSiBNb alloy, the permeability increases. On the other hand, coercive
force He' and core loss reduce remarkably by adding Cu.
Figure 4. 3 shows the Nb content dependence of magnetic properties in
Fe76.s-.Cu,Nb.Si'3.sBg alloys in optimum annealing condition (Yoshizawa and
Nanocrystalline Soft Magnetic Materials and Their Applications 127

12

10

<S 6 f=l kHz

Annealing conditions: 823 K, -3.6 ks

0.5 1.0 1.5 2.0


Cu comtent,x (at.%)

Figure 4. 2 Cu content x dependence of relative permeability mr in F7u-x Cu xNb3Sh35 8g


alloys annealed at 823 K for 3. 6 ks.

Figure 4. 3 Nb content dependence of magnetic properites in Fe765-a Cu,Nb aSi '35 8 9


alloys in optimum annealing condition.

Yamauchi 1989a, 1990a). The permeability increases and core loss decreases
with Nb content. Thus, the combined addition of Cu and Nb is a key point to
obtain high permeability over 1 x 10 5 , low core loss of about 280 kW/m 3 at
frequency f = 100 kHz, maximum magnetic induction 8 m = 0.2 T in FeSiB
system. The permeability is comparable with those of the amorphous Co-
based alloys and Supermalloy with extremely high permeability in soft
magnetic materials. Moreover, the saturation magnetic induction is over
1. 2 T, which is higher than that of Co-based amorphous alloys and
Supermalloy.
Figure 4. 4 shows the schematic figure of properties and composition
region in quasi-ternary (FeCu, Nb 3 ) SiB alloys, respectively. The composition
of zero magnetostriction exists in the composition region around 16 at. % Si
(Yoshizawa and Yamauchi, 1989b). The high saturation magnetic induction
128 Yoshihito Yoshizawa

over 1. 5 T is obtained in the composition of the Fe-rich region. The high


permeability is obtained in the composition region around 13.5 at. % Si, 9 at. % B.

30

60 L -_ _--,>~------:'-:--------:'----~40
Si 40 30 20 10 0 B
Si (at.%)

Figure 4. 4 Schematic figure of properties and composition region in quasi-ternary


(FeCul Nb3 ) SiB alloys.

Magnetic application products require controlling hysteresis loop shape for


magnetic materials. The static hysteresis loops (dc B-H loops) of the
nanocrystalline Fe73.5CulNb3Si,3.5Bg alloys for different heat treatments are
shown in Fig. 4. 5 (Yoshizawa and Yamauchi, 1989c). The rectangular B-H
loop (high remanence ratio), the round B-H loop (medium remanence ratio)
and the flat B-H loop (low remanence ratio) were obtained by annealing under
longitudinal field (magnetic field parallel to the magnetic path direction
(parallel to the ribbon axis, no-field, and transverse field (magnetic field
parallel to the ribbon width direction), respectively. A uniaxial magnetic
anisotropy is induced and the shape of the B-H loops can be controlled by
magnetic field annealing. The induced magnetic anisotropy change in field
annealing temperature and time can be controlled (Yoshizawa and Yamauchi,
199Gb) . The induced magnetic anisotropy of the nanocrystalline Fe-Cu-Nb-Si-B
alloy arises from the ordering of the Fe and Si atoms in the bcc Fe-Si phase.
The kinetics of induced anisotropy formation are much slower than those in the
amorphous state. The magnitude of the induced magnetic anisotropy is
determined by the fraction and Si content of the bcc Fe-Si phase and can
explain the induced magnetic anisotropy of a conventional Fe-Si alloy (Herzer,
1994a). The induced anisotropy in nanocrystalline Fe-Cu-Nb-Si-B alloys is
primarily determined by the silicon content and the fraction of bcc Fe-Si
grains. Since the available lattice sites for directional ordering decrease
simultaneously, the formation of induced anisotropy in bcc Fe-Si has been
proposed to arise from the directional ordering (Neel, 1954) of the nearest
Nanocrystalline Soft Magnetic Materials and Their Applications 129

B (T) H ma,=800 AIm


.-f=:;;;==----
1.0

Longitudinal field anneal

B (T) H max=800 AIm

No-field anneal
-----+nt--+----~H

Transverse field anneal

Figure 4. 5 Static hysteresis loops. (dc B-H loops) of the nanocrystalline


Fen. CUI Nb3 Si 13 . 8 9 alloys.

and next-nearest silicon-atom pairs. The permeability of nanocrystalline


FeCuNbSiB alloys is improved by controlling transverse anisotropy (Yoshizawa
et aI., 1995). By controlling transverse anisotropy, the flat B-H loop with high
relative initial permeability over 1 x lOs are obtained. The magnetic anisotropy is
also induced in the stress-annealed nanocrystalline Fens CUI Nb3Si 13 .S B9 alloys
(Kraus et aI., 1992). The anisotropy consists of creep-induced and magneto-
elastic components. By annealing under tensile stress, very strong anisotropy
is induced. In this case, anisotropy field H K = -4910 Aim of 2 orders higher
than the H K in the case of magnetic field annealing is induced by stress
annealing at 813 K under tensile stress a = 364 MPa. We should consider the
two-phase character of the nanocrystalline state with fine grains of bcc Fe-Si
phase and the amorphous matrix phase to understand the origin of creep-
induced anisotropy. It is thought that this anisotropy mainly originates from the
magneto-elastic anisotropy of the bcc Fe-Si grains due to tensile back stress exerted
by the anelastically deformed amorphous matrix (Herzer, 1994b).
Figures 4. 6 and 4. 7 give the frequency dependence of the relative
permeability and the core losses per core volume at B m = 0.2 T for the
nanocrystalline FensCulNb3Sil3.sB9 alloys annealed under or without magnetic
field (Yoshizawa and Yamauchi, 1991b, 1991c), respectively. The
Fens CUI Nb3Si I3 .S B9 alloy with low remanence ratio exhibits the highest
130 Yoshihito Yoshizawa

permeability and the lowest core loss in the high-frequency range. The
Fe73.sCu,Nb3Si,3.sBg alloy with high remanence ratio shows low core loss in the
high-frequency range as compared with conventional high remanence ratio
materials such as Fe-based amorphous and 50 wt. % Ni Permalloy, although
the core loss increases as compared with that of alloys annealed by other heat
treatment. This attractive high-frequency behavior is due to low hysteresis
loss, and low eddy current loss concerned with the thin ribbon thickness of
about 18 ~m and the high electrical resistivity of about 1. 2 ~Q m.

Fe73.5 Cu l Nb 3Si 13589


H=0.05A/m

102 '-:--'---'c..L.L.L.LllL.--'----'---L..L.LU..L'-:--'----'....>...LJ.LLU'-:--'--'---'-'-UJJ.J
10 10 1 102 103 104
Frequency (kHz)

Figure 4. 6 Frequency dependence of IJ' [complex permeability (real part) ] for


nanocrystalline Fe73.sCu,Nb3Si,3s8g alloys annealed under and without magnetic field.

Longitudinal field anneal

102
Frequency (kHz)

Figure 4.7 Frequency dependence of core losses at 8 m = 0.2 T for the nanocrystalline
Fe73.sCu,Nb3Si13s8g alloys annealed under and without magnetic field.

Figure 4. 8 shows the annealing temperature dependence of relative


permeability J1" saturation magnetostriction constant As and grain diameter D
Nanocrystalline Soft Magnetic Materials and Their Applications 131

for Fe73SCul Nb3Si,3sBg alloy (Yoshizawa and Yamauchi, 1991b, 1991c;


Yoshizawa, 1999). The high relative permeability over 1 x lOs was obtained
by anneal ing around 843 K for 3. 6 ks. This is due to the formation of the nano-
scale bcc Fe-Si grains from amorphous phase. On the other hand, when the
annealing temperature is higher than the optimum annealing temperature, the
formation of the compound phases such as Fe2 B with large magnetocrystall ine
anisotropy decreases drastically the permeability.
12
f=l kHz
10

b
~
8
, ,
J I
X
~ 6 Amorphous bcc+ Fe28,etc.
::t
4

0
700 750 800 850 900 950
Annealing temperature (K)

20
f'
0 15
x
~
10
Vl
...: 5
0
-5
700 750 800 850 900
Annealing temperature (K)

E
-8 10
Q
15

5
Fe73SCu INb3Si 13S 8 9

- -
o ,L.-.-'--_ _-'--_ _L.-._----J'---_-----'_ _---'
t 700 750 800 850 900 950
As-quenched Annealing temperature (K)

Figure 4. 8 Annealing temperature dependence of IJr' As and D for


FeJ3.5CulNb3Si13.5Bg alloy.

The Fe73SCu, Nb 3Si 13 . sBg alloy with the highest permeability annealed in
optimum annealing condition, circularly magnetized domain patterns are
observed by Lorenz microscopy. Since the domain wall width is about 5 times
as large as the area having a magnetic uniform anisotropy for the highest
permeability alloy, effects of the anisotropy field are averaged to zero,
resulting in free movement of the domain wall (Komoto et aI., 1990). It was
132 Yoshihito Yoshizawa

reported that the domain structure of the nanocrystalline Fe73.s CUI Nb3 Si 13 sB9
alloy observed by the Kerr effect resembles those of amorphous alloys
(Schafer et a!. , 1991). At temperatures over 320 C, the domain structure
changes because the amorphous matrix phase transformation to the
paramagnetic state decreases the interaction among ferromagnetic crystal
grains. The similar domain pattern is also observed in the overannealed material.
The nanocrystalline soft magnetic materials are characterized by the fact
that the Curie temperature of the bcc phase is higher than that of the
amorphous ferromagnetic materials or soft ferrite materials. The temperature
dependence of the magnetic properties is significant for applications. The
temperature dependence of relative permeability and core losses per core
weight are shown in Figs. 4. 9 and 4. 10, respectively. The nanocrystall ine
FeCuNbSiB alloys exhibited extremely good magnetic stability against the
temperature in the temperature range from 223 K ( - 50 C) to 423 K ( + 150 C).

105~
Fe-Cu-Nb-Si-B(Transverse field anneal)
Co-based amorphous alloy'
....... -_-e'~---- __
~~-*-----------~~"
_..............
.~~.~~~.~~~.~.

_........._A- -+--+--+- ..
\
\

... Fe-Cu-Nb-Si-B(No-field anneal) )


3
10 Mn-Zn ferrite \
\
j=20kHz \
\
10~50 0 SO 100 150
Temperature CC )
Figure 4.9 Temperature dependence of relative permeability for nanocrystalline
FeCuNbSiB alloys, an amorphous Co-based alloy and Mn-Zn ferrite.

16

j=20 kHz
Bm=0.2 T
12

4 Fe-Cu-Nb-Si-B (No-field anneal)

o Fe-Cu-Nb-Si-B(Transverse field anneal)


-SO 0 SO 100 150
Temperature (C )

Figure 4. 10 Temperature dependence of core losses per core weight for nanocrystalline
FeCuNbSiB alloys and Mn-Zn ferrite.
Nanocrystalline Soft Magnetic Materials and Their Applications 133

The change in permeability and core loss was small in the temperature range
from 223 K ( - 50 C) to 423 K (+ 150 C). Therefore, this behavior of
FeCuNbSiB alloys can be used in a wide temperature range in comparison with
ferrite materials.
Nanocrystalline soft magnetic materials exhibit a small magnetic change
against time as compared with amorphous soft magnetic materials. The
disaccommodation (D. A.) decreases remarkably by nanocrystallizing the
amorphous Fe73. 5CU, Nb3 Si ,6 . 5B6 alloy (Z6veta et aI., 1990). The magnetic
temperature stability is one of the merits of the nanocrystalline soft magnetic
alloys for applications.

4. 2. 2 Microstructure
The nanocrystalline FeCuNbSiB alloy consists of ultrafine bcc Fe-Sicrystal
grains of about 10 nm in average grain diameter, residual amorphous matrix
phase, and fcc Cu(Fe) grains about 5 nm (Yoshizawa and Yamauch, 1989a;
Hono et aI., 1992).
Figure 4. 11 shows the microstructure of Fe74.5-x CU x Nb3 Si 13 . 5 B9 alloys
annealed at 823 K for 3.6 ks. The grain size decreases remarkably and the
grains become more homogeneous by adding Cu. This explains the Cu
dependence of soft magnetic properties.
~

(c) x=LO (d) x=L5

Figure 4. 11 Microstructure of Fe74.5-. Cu. Nb3 Si,3.S 8 9 alloys annealed at 823 K for 3.6 ks.
134 Yoshihito Yoshizawa

Figure 4. 12 shows the microstructure of Fe76.S-a Cu, Nb a Si 13 . s B9 alloys.


The grain size decreases with Nb content. This suggests that Nb suppresses
the grain growth.

(a) a = 0(693 K,3.6 ks) (b) a = 1(773 K,3.6 ks)

(e) a=2(823 K,3.6ks) (d) a = 5(873 K,3.6ks)

Figure 4. 12 Microstructure of Fe76.S-a CUI Nb a Si l3 . s Bg alloys.

Thus, the combined addition of Cu and Nb is effective in reducing grain


size. Yoshizawa and Yamauchi (1989a, 1990a) suggested that the Cu
enhances the formation of bcc Fe(Si) crystal nuclei, and the Nb stabilizes an
amorphous matrix phase and suppresses the grain growth of bcc Fe(Si) crystal
grains, so that the homogeneous and ultrafine grains would be formed. The
magnetic properties are strongly correlated to the microstructural features.
However, Cu is not necessary for reducing grain size in some alloy systems
such as FeZrB alloys (Zhang et aI., 1996).
In the nanocrystalline FeCuNbSiB system alloys, Si and B content affect
the microstructure and the magnetic properties (Yoshizawa and Yamauchi,
1989b; Ueda et al., 1994). The lattice parameter of ex-FeSi on annealing
shows a strong dependence on the concentration of silicon. The minimum value
of the lattice parameter of ex-Fe is obtained at an optimum annealing temperature.
Figure 4. 13 shows X-ray diffraction patterns of Fe73.SCUl Nb3 Si I3 . SB9 alloy
annealed at different annealing temperatures.
The volume fraction of the bcc Fe-Si phase increases with annealing
temperature by annealing over crystallization temperature, and the highest
permeability is obtained in the annealing condition where the main phase is bcc
Fe-Si phase. By an X-ray diffraction and M6ssbauer effect, it becomes clear
that the bcc Fe-Si phase precipitated in FeCuNbSiB alloys with high Si content
Nanocrystalline Soft Magnetic Materials and Their Applications 135

bee Fe(Si)

Fe73.5CU I Nb )Si 13.589


Cu-K"

bee Fe(Si)

J+'11-Jf:-.-'::"!';~~~~~~~fp.~~973K,3.6 ks
923 K,3.6 ks
873 K,3.6 ks
,",.:z:::;:~:::;::;:::::;:2:::::="?1
, 773 823 K,3.6
K,3.6 ks ks
723 K,3.6 ks
60 70 80 90673 K,3.6 ks
Volume fraction

Figure 4. 13 X-ray diffraction patterns of Fens CUI Nb3 Si,35 89 alloy annealed at different
annealing temperatures.

has the D03 super lattice.


When the annealing temperature is much higher than optimum annealing
temperature, in spite of keeping the bcc Fe-Si grains small, the drastic
permeability decrease occurred. This behavior is due to precipitating borides
such as Fez B with high magnetocrystalline anisotropy.
The many works on the effects of the substitution of the M elements (M:
V, Cr, Mo, Ta, W, etc.) for the Nb element on the microstructure and
magnetic properties in the FeCuMSiB alloys and the effects of the addition of
the M elements on the microstructure and magnetic properties have been
carried out (Yoshizawa and Yamauchi, 1991d; Duhaj et aI., 1991; Brovko
et aI., 1993; Muller and Mattern, 1994; Muller et aI., 1993, 1996).
Figure 4. 14 shows the grain diameter D of Fe73.sCulM3Si13.sBg alloys (M:
Ti, V, Cr, Mn, Zr, Nb, Mo, Hf, Ta, W)(Yoshizawa and Yamauchi, 1991d).
60
Fe73.5 Cu \M)Si 13.589
50

40

~ 30
Q
20

10

Ti V Cr Mn Zr Nb Mo Hf Ta W
M
Figure 4. 14 Grain diameter 0 of FEry35 CUI M3 Si l3 . s8 9 alloys(M: Ti, V,
Cr,Mn,Zr,Nb,Mo,Hf, Ta,W).
136 Yoshihito Yoshizawa

The average grain diameter is below 20 nm in the case of M = Nb, Ta, Zr, Hf,
Mo, W. However, the grain size in the case of M=Cr, Mn is larger than that
in the case of M = Nb, Ta Zr, Hf, Mo, W.
Hono et al. found that one phase is the bcc-Fe-Si solid solution in the
Fe73.5 CUI Nb3 Si,3.5 8 9 alloy, and the other phase was a remaining Nb and 8
enriched amorphous phase by atom probe field ion microscopy study (1991,
1993). In addition, Hono et al. reported that the Cu clusters are formed in the
early stage of the annealing from the analysis results by atom probe field ion
microscopy on the microstructure and crystallization process of
Fe73.5 CUI Nb3Si13.589 alloy (1992). They reported that a separate nanobeam
electron diffraction study reveals that the Cu enriched particles are fcc-Cu in
optimum annealing condition. The precipitated bee-Fe phase contains 20 at. %
Si, but hardly contains 8 and Nb in optimum annealing condition. Most of the
Nb and the 8 are contained in a residual amorphous matrix phase. The
amorphous matrix phase of FeNbSi8 is stabilized by the increase of the Nb and
the 8 content.
Hono et al. confirmed that the non-magnetic fcc Cue-Fe) grains of about
5 nm in grain diameter exist in the nanocrystalline Fe73.5 CUI Nb3Si 13 .58 9 alloy
directly. From this direct analysis by atom probe field ion microscopy
(APFIM), they proposed the crystallization process and refinement mechanism
(Hono et al. , 1992). They made it obvious that the formation of bee nuclei by
forming Cu clusters and the suppressing the grain growth of the bcc-Fe-Si
grains by Nb and 8 contributes to the refinement of the grains. Sakurai et al.
reported that a structure similar to fcc Cu exists in the early stage of heat
treatment from the analysis by X-ray absorption fine structure(XAFS) (Sakurai
et aI., 1994).
Reoently, Hono et al. analyzed the nano-structure of the Fe,3.5 CuI ~ Si,3.5 8g
alloy by using a three-dimensional atom probe (3DAP) and a proposed
crystallization process (Hono et aI., 1999). They reported that the 3DAP
results clearly show that Cu atom clusters are present in the amorphous state
after annealing below the crystallization temperature. The density of these
clusters is about 1 x 1024 m- 3 , which is comparable to that of the bee-Fe grains
in the optimum nanocrystalline microstructure. In the early stage of
crystallization, the direct contact of the Cu clusters with bee-Fe nanocrystals
suggests that each bee-Fe primary particle is heterogeneously nucleated at the
site of Cu clusters.

4. 2. 3 Origin of Magnetic Softness

4.2.3.1 Grain Diameter


In Permalloy thin films, the theory to explain the soft magnetic properties when
the crystal grains become extremely fine, was proposed (Hoffman, 1964).
Nanocrystalline Soft Magnetic Materials and Their Applications 137

However, this theory was not believed as the basic principle for the
development of the soft magnetic materials with extremely high permeability,
because the soft magnetic properties of the Permalloy thin films are inferior
more remarkably than those of the bulk Permalloy with large grains. However,
the discovery of the nanocrystalline FeCuNbSiB materials with extremely high
permeability has generally recognized that the decrease of grain size to nano-
scale leads to a magnetic softening for the crystalline magnetic materials.
Figure 4. 15 shows theoretical estimate of the average anisotropy constant
<K > for randomly oriented o:-FeSi grains as a function of grain diameter D
(Herzer, 1989). Herzer suggested that magnetic softening in nanocrystalline
materials could be theoretically explained by a random anisotropy model
(Alben et aI., 1978). Herzer reported that the relation where the coercive
force He of the nanocrystalline materials is proportional to the 6th power of
grain diameter D, and the permeability varies as the minus 6th power of grain
diameter D is in good agreement with the experimental data (Herzer, 1990).
10

o 10 20 30
Grain diameter, D(nm)

Figure 4. 15 Theoretical estimate of the average anisotropy <K) for randomly oriented
cx-FeSi grains as a function of grain diameter D (Herzer, 1989) .

Although the potential in the magnetization process of a single crystal


with magnetocrystalline anisotropy K 1 and saturation magnetization M s is
given by K, , this is given by K 1 1..jN if grain diameter D decreases below a
domain wall width. Where the N is the number of the crystal grains included in
the area (the exchange area) where the magnetization is almost uniform.
When the grains are extremely fine and homogeneous, ferromagnetic exchange
length Lex with (D<L ex ) related to Nand N is given by N= (L ex ID)3. The
Lex is given by Lex = (AI <K 1/2, which is determined by the magnetic
anisotropy of magnetization within Lex and exchange energy (exchange stiffness
constant A). The effective magnetic anisotropy <K > for nano-scale grains is
determined by the mean fluctuation amplitude of the anisotropy energy.
Accordingly, we find the next formula (Herzer, 1989, 1990).

(4.1)
138 Yoshihito Yoshizawa

This means that the <K) decreases with decreasing grain size. This holds
as long as the grain size D is smaller than the exchange length Lex. If
coercivity He and initial permeability Pi are related to <K) using the results for
coherent magnetization rotation, we find (Herzer, 1990):

(4.2)

(4.3)

where Po is permeability of free space.


Hence, He and PI should be sensitively affected by grain size. These
formulas mean that the He and the PI are proportional to the 6th power of D
and the minus 6th power of D, respectively. The same result can be derived
assuming domain wall pinning as the prevailing magnetization mechanism.
As shown Fig. 4. 16, the experimental data plots for He and PI are both
compatible with the expected dependence (He is proportional to the 6th power
of D, PI is proportional to the 6th power of D) (Herzer, 1990). Also, if
induced magnetic anisotropy constant K u dominates over < K ), He varies
proportionally to the 3rd power of D instead of the 6th power of D. A D 3
dependence of He is observed for some nanocrystalline samples. Suzuki et al.
reported that the extended random anisotropy model predicts a D 3 dependence
of He when uniaxial anisotropies are larger than <K) (1998).
\00
,.-. +: Amorphous Fe-base

10 ,.' \ . .,
,
x: Amorphous Co-base
' , D - I . : Nanocrystalline Fe-CUo.INb3(SiBb.5
E f ' 0: Fe-Si6.5wt.%

~
,
',0
0'
0: 50 Ni - Fe
~: Pennalloy
:z; D6
, ,
0' 0
S : 0',

. ,
'>
.~
0.\ ~,
, r:J\ o
(l)
o
u 0.0\
+ t ~ ~
",
*----, ~~ .
0.00\1.......---'--------,--'------"..-'--
1 nm 1 Jlm \ mm
Grain size, D

Figure 4.16 Grain size and coercivity He for various soft magnetic materials (Herzer, 1990).

On the other hand, if the grain size exceeds the domain wall width (D >
TrL ex ) , the magnetization process is determined by domain wall pinning at the
grain boundaries. He and permeability PI are given by:

JA. K
He = Pe ---=-----.-....:..1
MsD
(4.4)

M2 D
PI = P~ _--;:s::;=~=- (4.5)
Po J A K 1
Nanocrystalline Soft Magnetic Materials and Their Applications 139

with pre-factors being typically Pe = 6 (3) and p" = O. 1(0.2) for 180 (90)
domain wall in an assembly of randomly oriented cubic grains.
The measured data for He and PI of the conventional crystalline bulk soft
magnetic materials such as Permalloy and silicon steels are proportional to
0- 1 , and 0, respectively (Herzer, 1990). This means that the soft magnetic
properties in conventional crystalline bulk materials are improved with
increasing grain diameter D.
4.2.3.2 Intrinsic Magnetocrystalline Anisotropy
As expected from Eqs. (4.1) - (4.3) (Herzer, 1990), the decreasing intrinsic
magnetocrystalline anisotropy K 1 leads to the decrease of He and the increase
of PI due to the decrease of the mean magnetic anisotropy < K ). In
nanocrystalline FeCuNbSiB system alloys, bcc Fe-Si crystal grains are
dispersed in an amorphous matrix. It is well known that the K 1 of the bcc Fe-Si
decreases with Si content (Trasov, 1939). Therefore, the soft magnetic
properties of the nanocrystalline FeCuNbSiB alloys are affected by Si content.
The maximum value of initial permeability in nanocrystalline FeCuNbSiB alloy
is obtained around 13. 5 at. % Si and 9 at. % B where magnetostriction is not
zero, but slightly positive. Herzer reported that this is related to the
contribution of magnetocrystalline anisotropy that increases with Si content due
to a corresponding increase of the average grain size (Herzer, 1992). The
composition of the highest permeability agrees with the composition of the
minimum of < K) which is determined by K l ' 0 of bcc Fe-Si and
magnetoelastic anisotropy constant K 01 related to the balance in
magnetostriction among the bcc Fe-Si grains and residual amorphous matrix.
In addition, the Fe-Cu-Nb-Si-B alloys form compound phases such as Fe2 B with
higher magnetocrystalline anisotropy than bcc Fe-Si phase in the annealing
temperature over optimum annealing temperature (Yoshizawa and Yamauchi,
1991 b, 1993). The soft magnetic properties deteriorate drastically, even
though the grains of the bcc Fe-Si stay small. This behavior suggests that the
soft magnetic properties are also affected by the amplitude of the intrinsic
magnetocrystalline anisotropy of the crystalline phases. This means that the
good soft magnetic properties in the nanocrystalline magnetic materials are
realized by forming the nano-scale bcc Fe-Si crystal grains with small intrinsic
magnetocrystalline anisotropy and no compound phases with large
magnetocrystalline anisotropy. The soft magnetic properties in the
nanocrystalline FeCuNbSiB alloys deteriorate if the compound phases with
large magnetocrystalline anisotropy are formed. The magnetic properties
change drastically and the magnetization occurs by rotational processes when
the compound phases are formed (Reininger et aI., 1992). They showed that
the corresponding change in the magnetization process is attributed to
increased pinning of domain walls due to the precipitation of Fe-B compound
phases.
It is significant for excellent soft magnetic properties in the nanocrystalline
140 Yoshihito Yoshizawa

soft magnetic materials to precipitate the ferromagnetic grains with small K 1


by choosing the annealing condition and composition as mentioned above.

4.2.3.3 Magnetostriction
We can consider two ferromagnetic phases for the nanocrystalline FeCuNbSiB
alloys, although, strictly speaking, the non-magnetic fcc Cu phase exists. The
Fe-based amorphous phase with large positive magnetostriction and the bcc
Fe-Si crystal grains dispersed in the Fe-based amorphous matrix phase in the
nanocrystall ine FeCuNbSiB alloys, approximately. The total saturation
magnetostriction As is given by
(4.6)
where the A~esl, V Fesl and Ar' denote the saturation magnetostriction of the bcc
Fe-Si phase, the volume fraction of bcc Fe-Si phase, and the saturation
magnetostriction of amorphous matrix phase, respectively (Herzer, 1991,
1993,1996). The A~esl of FeSi is-5x 10- 6 (Fe-20 at.%Sj) and the Ar' is
about 10 x 10- 6 -25 X 10- 6 of the positive value. As understood from Eq. (4. 6),
the increasing VFeSi by crystall izing the amorphous alloys leads to decreasing As.
When heat treatment temperature is higher than crystallization temperature, the As
decreases and the permeability increases drastically as shown in Fig. 4.8. This
behavior suggests that the decrease of the A s contributes to improving permeability.
Magnetoelastic anisotropy constant K el is given by
3 (4.7)
Kel="2IAsoal

where the As and the a denote saturation magnetostriction constant and


internal stress, respectively. If the a is constant, we can get the relation
where He is proportional to the absolute value of As and where IJr is
proportional to the absolute value of As' .
Figure 4. 17 shows the relationship between IJr and absolute value of As'
for Fen5 CUI Nb3 Si 13 . 5 B9 alloy annealed in different annealing temperature.
12 I ,
I ._
/=1 kHz
10 I I
: I -
-: Fe73SCuINb3SiI3.5B9
I : Fe7JCulNb3SilsBs

x 6
'-'
,: -l\ -
, \..1. -I
::t ~ \ s
4 1\ ,/
I -..1. -I ~
2/ s "-
...
ol..--_-'-_---'_ _-'---_---'_ _--'-_---'
_--~ ,..-
---------
-5 0 5 10 15 20 25
..1.s(X 10-6)

Figure 4. 17 Relationship between IJr and absolute value of As 1 for F~3.5 CUI Nb3 Si 13 . 5 Bg
alloy annealed in different annealing temperatures.
Nanocrystalline Soft Magnetic Materials and Their Applications 141

The data of fJ r is almost proportional to A;-'. This data indicates that the
increasing fJr in the nanocrystalline state results from the decrease of As due to
the precipitation of bcc Fe-Si phase with low magnetostriction by crystallizing
Fe-based amorphous alloys.
Figure 4. 18 shows silicon content y dependence of saturation
magnetostriction constant As, grain diameter 0, lattice parameter a in the
Fe73.5 CUt Nb3SiyB22.5-y alloys (Yoshizawa, 1999). The As decreases with Si
content, and crosses almost zero around 16 at. % Si.
12
10 /=1 kHz
~

b 8 No-field anneal
x
~ 6
::t 4
2

0 5 10 15 20

f'
-x
0
--. Si content, y (at.%)

'.
~

<'<
Vl
'H
-5 0 5 10
:~ IS 20
Si content, y (at.%)

16
E
oS 12
Fe735Cu,Nb3Si)322.5_y
I
J
Q Optimum annealing condition
8
0 5 10 15 20
Si content, y (at.%)
0.286

f02"f
" 0.284
0.283
0
I

5
~ 10
I

15
I I

20
Si content, y (at.%)

Figure 4. 18 Silicon content y dependence of saturation magnetostriction constant As,


grain diameter D,lattice parameter Q in the Fe73.sCu,Nb3SiyB22.s-yalloys.

The As shows the negative value over the Si concentration of zero


magnetostriction. Herzer (1991) reported that the As of the Fe73.5 Cut NbJ Siy B.!2.5- y
alloys is explained by the Si content dependence of volume fraction and
magnetostriction of the bcc Fe-Si phase and residual amorphous matrix phase
with positive A s as understood from Eq. (4. 6) .
Figure 4. 19 shows the relationship between /-Ir and saturation
magnetostriction constant A s of quasi-ternary (Fe-Cut -Nb3) -Si-B alloys. The
maximum permeability is not found for As = 0 composition but at a lower silicon
concentration with slightly positive As. This behavior suggests that it is
142 Yoshihito Yoshizawa

significant for the magnetic softening to reduce the total magnetostriction As;
however, the soft magnetic properties in the nanocrystalline materials cannot
be completely explained by only the total magnetostriction As.
12 , ,
,I
, (Fe-Cu l-Nb3)-Si-B
10
,, ~.
I Optimum annealing condition

8 ,, ....~"" ..
f=l kHz

,'" . . \
\
x
~

:t 4
I

/\ . \
\
\
~ A-I
/ _A-I
s '" ......IS
2 ~
---- . .... _--
........
~--
0'-----''------'------'-----'----'
-5 o 5 10 15 20
As (X 10-6)

Figure 4. 19 Relationship between IJ, and saturation magnetostriction constant As of


quasi-ternary(Fe-Cu,-Nb3 )-Si-B alloys.

4. 2. 3. 4 Influence of Grain Boundary Phases


As mentioned before, the nanocrystalline FeCuNbSiB alloys consist of three
phases, i. e., ferromagnetic bcc Fe-Si grains, non-magnetic fcc-Cu phase,
and ferromagnetic Nb and B enrich Fe-based amorphous matrix phase. Each
bcc Fe-Si crystal grain interacts magnetically through the ferromagnetic
amorphous matrix phase.
Accordingly, the soft magnetic properties deteriorate due to the decrease
of the exchange interaction if the magnetization of the amorphous matrix phase
decreases. In the extreme case, for example, when the residual amorphous
matrix phase becomes the non-magnetic phase, drastic increase in coercive
force and drastic decrease in permeability occurs (Herzer, 1989). Therefore,
if the grain boundary phase shows the extremely low Curie temperature T c and
low magnetization, it is difficult to use for a practical use, even if the soft
magnetic properties are excellent at room temperature.
Figure 4.20 shows Nb content as dependence of relative permeability IJ"
grain diameter D of bcc Fe-Si grains and Curie temperature T c of the
amorphous matrix phase in Fe76.S-aCulNbaSi13.sBg alloy (Yoshizawa and
Yamauchi, 1991b). Maximum value of relative permeability is found from
3 at. % to 5 at. % Nb. However, although the D decreases with increasing Nb
content, relative permeability, IJ, shows maximum value in the composition
range of from 3 at. % Nb to 5 at. % Nb. The IJ, decreases when Nb content
becomes higher than 5 at. % Nb. The T c of the remaining amorphous matrix
phase decreases with increasing Nb content because the Nb is partitioned in
the amorphous matrix phase. Accordingly, the magnetization of the amorphous
matrix phase is estimated to decrease with increasing Nb content. Hence, the
Nanocrystalline Soft Magnetic Materials and Their Applications 143

large amount of Nb addition weakens the exchange interaction among the


ferromagnetic bcc Fe-Si grains, even though the grain size decreases with
increasing Nb content.
14
,-,. 12
b 10
6 8
::l: 6
4
H=0.05A/m
2

0 2 3 4 5 6 7
Nb content, a (at.%)

Fe76.5_aCU I NbaSi 13.5 8 9


E 20
5
Q
10

0 2 3 4 5 6 7
Nb content, a (at.%)

900 Residual amorphous


matrix phase
800
g 700
~ 600
500
400
0 2 3 4 5 6 7
Nb content, a (at.%)

Figure 4. 20 Nb content dependence of relative permeability IJr ,grain diameter D of bcc-


Fe-Si grains and Curie temperature T c of the amorphous matrix phase in
Fe76.s-aCu,Nb aSi,3.s8g alloy.

Herzer (1995) reported the extension of the random anisotropy model to


two-phase systems. Suzuki and Cadogan (1998) reported two-phase random
anisotropy model with two local exchange stiffness constants. The
experimental results are quite well explained by the proposed model.

4. 3 Applications

Nanocrystall ine soft ferromagnets have features as follows:


(1) High saturation magnetic induction (comparable to Fe-based
amorphous alloys) ;
144 Yoshihito Yoshizawa

(2) High permeability (comparable to Co-based amorphous alloys or


Supermalloy) ;
(3) Shape control of hysteresis loops (many applications) ;
(4) Low core loss;
(5) Low magnetostriction (zero magnetostriction is possible in some
case) ;
(6) Good temperature stability (in the range of from-50 C to 150 C);
(7) Small magnetic property change against time (in comparison with Co-
based amorphous alloys);
(8) Good high-frequency characteristics.
Figure 4.21 shows the features and typical applications for nanocrystalline
FeCuNbSiB alloys (Yoshizawa, 1999). Typical applications in the
nanocrystalline alloys are shown as follows:
(1) Coil parts for noise measures of personal computers (PC), machine
tools, lasers and power supplies;
(2) Choke coils or transformers for inverters and power supplies;
(3) Thin parts for communications such as pulse transformers used for
interfaces (PC cards for note PC) of integrated services digital networks
(ISDN) ;
(4) Large magnetic core components used for pulsed power applications
such as linear induction accelerators, radio frequency (RF) cavity for
synchrotrons, magnetic pulse compression circuits for XeCI (excima) lasers,
or accelerators for medical use;

( High B s )
~ C Good temperature stability -.:J
Noise filters
Common-mode choke coils
Pulse transformers
Current transformers
Sensors
(High Bs ) Electromagnetic shielding ( High Bs )
L?WAs \J. '-Wave absorber "" ( Low core loss.J
:\
C
( Low core lOss)
( HighEr )
Choke coils
Power transformers Noise absorbers
Active fi Iters Bead cores
Cut cores Magnetic switches
Dust cores Saturable reactors

Figure 4.21 Features and typical applications for nanocrystalline FeCuNbSiB alloys.
Nanocrystalline Soft Magnetic Materials and Their Applications 145

(5) Magnetic parts used for measuring or detecting currents in a circuit,


such as breakers and current sensors;
(6) Magnetic shielding.

4.3.1 Parts for Noise Measures


Common-mode choke coils, noise absorbers, and bead cores are commer-
cialized as application parts for noise measures.
4. 3. 1. 1 Common-mode Choke Coils
Common-mode choke coils are used to attenuate common-mode noises, which
are conducted to power lines from machine tools, power supplies for lasers or
high-voltage pulse noises, which are conducted to electronic equipment from a
power line. The common-mode choke coils of the nanocrystalline Fe-Cu-Nb-Si-B
alloys exhibit excellent noise attenuation characteristics due to their high
saturation magnetic flux density and high permeability (Yoshizawa et al. ,
1988b).
Figure 4.22 shows the external view of common-mode choke coils using
the nanocrystalline FeCuNbSiB alloy cores. The common-mode choke cores
for noise filters were first put to practical use as magnetic parts of
nanocrystalline Fe-Cu-Nb-Si-B alloys. Figure 4.23 shows the frequency
dependence of attenuation for common-mode choke coils. Figure 4. 24 gives
pulse attenuation characteristics for common-mode choke coils (Yoshizawa et
aI., 1988b). The common-mode choke coils of nanocrystalline soft magnetic
materials can realize about half size for the common-mode choke coils of
ferrite cores and large noise attenuation.

Figure 4.22 External view of common-mode choke coils using the


nanocrystalline FeCuNbSiB alloy cores.

The nanocrystalline ferromagnet cores are also used as zero phase


reactors (radio frequency reactors). The electromagnetic compatibility(EMC)
measure parts of the nanocrystalline soft magnetic materials show high
performance for power electric devices, and especially components used in
the frequency range below several MHz.
146 Yoshihito Yoshizawa

50
Common mode
Core size: 20, 10, 6.5
40 12 turns

m 30

_....- --------.,
~
.Q

J ///;:~//~'fi~~~~:;)
20
/ FeCuNbSiB(nano)
./ / (Transverse field anneal)
./ - - - Mn-Zn ferrite

10-\ 10 10 1
Frequency (MHz)

Figure 4.23 Frequency dependence of attenuation for common-mode choke coils.

200
Pulse width: 800 ns Core size: 20, I0,6.5
- - FeCuNbSiB(nano)(No-field anneal)
~ .......... FeCuNbSiB(nano)(Transverse field anneal)
..,.0" 150 - - - Fe-based amorphous alloy !

~
- - Mo-Zo '''rij !
"0
~ 100
& / /!
12 turns / / !
'"
///
Ul
:;
0..
::; 50 /
&
o'"
/ ./
....
....
/ .;
/
....
.'

o 200 300 400 500 600


Input pulse peak voltage, Vin(V)

Figure 4.24 Pulse attenuation characteristics for common-mode choke coils.

4. 3. 1. 2 Bead Cores. noise Absorbers


Bead cores and noise absorbers are used to control the recovery currents of
diodes. to control the generation of noises, to control the spike currents of
switching circuits, to prevent ringing, to protect shotky barrier diode from
reverse voltage, and so on.
Figure 4.25 shows an example of a circuit using bead cores of a nanocrystalline
FeO.JNbSiB material and waveforms of every part before and after using bead cores
( Yoshizawa et al.. 1997). By using nanocrystall ine soft magnetic material
"FINEMEr" beads, the recovery currents and output noises reduce substantially,
because of their high remanence ratio and high permeability.
Nanocrystalline Soft Magnetic Materials and Their Applications 147

1"F1NEMET" beads I

-
:5 ------ ~~ 1 =
~VI2!

CD ~"""~.
0-

ac
~
'u....
-< >-
input
1::
Q) -< >- =e
-< ~ >-

------ ~ ~
o-~J-J

Unmeasures Measures
Recovery wave forms
Diode voltage (top)
10.0 V/div.
Diode current (bottom)
5.0 Ndiv.
I s/div.

t= ~
Recovery wave forms
(Expansion)
Diode currents
500mNdiv.
0.1 s/div.

tt ~
Output noise wave forms
Output voltage
0.2 V/div.
I s/div.

Used diodes: LLD (FRO) trr.=IOO ns

Figure 4. 25 Example of a circuit using bead cores of a nanocrystalline FeCuNbSiB


material and waveforms of every part before and after using bead cores.

4. 3. 2 Magnetic Parts for Communications


Pulse transformers used for PC cards for integrated services digital networks
(ISDN) require a thin model and high inductance at the frequency of about
20 kHz. Hence, the high permeability is required for the core materials in this
frequency range. In addition, good temperature stability is required for the
pulse transformers.
Figure 4.26 shows an appearance of a pulse transformer using
nanocrystalline Fe-Cu-Nb-Si-B alloy cores for PC cards (3 mm in thickness) for
ISDN. An impedance characteristic of a pulse transformer for ISDN is shown in
Fig. 4. 27. The pulse transformer with 2.9 mm in thickness of the
nanocrystalline soft ferromagnets satisfies the standard value sufficiently. If we
use a nanocrystalline core as a pulse transformer for ISDN, the size of the
transformer becomes small, and the withstand voltage becomes high as
compared with pulse transformer of ferrite cores. In addition, a good
temperature stability and compact size can be achieved (Yoshizawa et al. , 1997).
148 Yoshihito Yoshizawa

Figure 4. 26 Appearance of a pulse transformer using nanocrystalline FeCuNbSiB alloy


cores for PC cards (3 mm in thicknessHor ISDN.

102 L-..l..-_---' --'- --'-_ _- - - '

10 10 1 102 103 104


Frequency (kHz)

Figure 4. 27 Example of impedance characteristic of pulse transformer for ISDN.

4.3.3 Parts for Pulsed Power

Magnetic switch cores are used for pulsed power applications such as the
linear induction accelerators or power supplies in pulse compression circuits
for the excimer lasers. These applications are required to compress high
electromagnetic energy instantaneously.
4.3.3.1 Magnetic Switch Cores for Excimer Lasers
Nanocrystalline FeCuNbSiB alloy cores annealed under a magnetic field
parallel to a magnetic path direction of cores are used for pulsed power
applications. because the down-sizing and low core loss are materialized.
This is due to higher remanence ratio. higher magnetic induction and lower
core loss in the short pulse width range than those of Fe-based amorphous
Nanocrystalline Soft Magnetic Materials and Their Applications 149

cores, and higher full magnetic swing than those of Co-based amorphous cores
(Sajiki et ai., 1995).
Figure 4.28 shows a circuit example of an excimer laser using a magnetic
switch core. A magnetic switch core converts an input broad pulse current into
a sharp and high output current. As a result, a laser beam is generated from a
diode. Table 4.2 gives the comparison of magnetic switch cores. By using
nanocrystall ine cores, compact cores with low loss can be real ized.

Figure 4. 28 A circuit example of an excimer laser using a magnetic switch core.

Table 4.2 Comparison of magnetics switch cores.

Fe-based amorphous Co-based amorphous


Materials FINEMEr
alloy alloy

Reset magnetization
3.55 7.1 3.55
forceCA!m)

Core volume ratio 1. 35 1.0 7.5


Core loss ratio 0.55 1.0 0.55

4. 3. 3. 2 Magnetic Switch Cores for Accelerators


The nanocrystalline FeCuNbSiB alloy cores are used as the cores for linear
induction accelerators, which are used to accelerate charged particles
(Nakajima et ai., 1993; Watanabe et ai., 1998). Recently, nanocrystalline
alloy cores are also used for RF cavities (Saito et al., 1998; Mori et al. ,
1998; Koba et ai., 1999), which can be appl ied to a heavy ion synchrotron
and a proton synchrotron for nuclear physics, as well as for medical proton
synchrotron. The untuned RF cavities of the nanocrystalline Fe-Cu-Nb-Si-B
alloy cores can realize both a wide operating frequency range and a high
stable accelerating gap voltage.

4. 3. 4 Parts for Power Supplies

4. 3. 4. 1 Saturable Cores
Saturable cores are used for controlling output voltage in switched-mode power
supplies. The cores of nanocrystalline soft magnetic alloys with high
150 Yoshihito Yoshizawa

remanence ratio fabricated by magnetic field annealing are used for saturable
cores of switched-mode power supplies of electronic equipment such as
personal computers (Nakajima et al. , 1999), because they show the excellent
constant current flux reset (CCFR) characteristics comparable to saturable Co-
based amorphous cores.
Figure 4.29 shows constant current flux reset (CCFR) characteristics for
the various saturable cores. The nanocrystalline soft magnetic core shows
good CCFR characteristics comparable to Co-based amorphous cores.
2.5
/=50 kHz
2.0 . FeCuNbSiB(nano)
... : Co-based AM(Type I)
.: Co-based AM(Type 2)
1.5 "': 80% Ni-Fe
E .: 50% Ni-Fe
<:l::l 1.0

0.5

oL---..L_--l...._::L:!=:t:=iI==Ijb"",...,.l~.".
-70 -60 -50 -40 -30 -20 -10 0
R(A/m)

Figure 4.29 Constant current flux reset(CCFR)characteristics for the various


saturable cores.

4.3.4.2 Transformer Cores for Inverters


Silicon steels, Fe-based amorphous materials, etc. with high magnetic flux
swing have been widely used as core materials for high power capacity
inverters operated in the frequency range of several 10kHz or less.
The Fe-based amorphous alloys exhibit relatively low core loss.
However, although they exhibit a low core loss before molding, the large
magnetostriction of about 20 x 10- 6 in Fe-based amorphous alloys makes the
core loss increase, if they are molded by a resin to fabricate cut cores for
inverter transformers. The cut cores of the nanocrystalline soft ferromagnets
with almost zero magnetostriction show one order lower core loss than Fe-
based amorphous cut cores, because the increase of the core loss after
molding is small due to almost zero magnetostriction. Hence, the cut cores of
nanocrystalline soft ferromagnets are available for the inverter transformers
that are operated in the frequency range of several 10 kHz or less (Fukunaga
et al. , 1990).
Figure 4. 30 shows the frequency dependence of core losses for various
cut cores. The cut core of the nanocrystalline soft magnetic alloy exhibits
lower core loss than the conventional cut cores.
The nanocrystalline Fe-Cu-Nb-Si-B toroidal cores annealed under
Nanocrystalline Soft Magnetic Materials and Their Applications 151

102

Oil
.-'<

~ \01
E
0...0

\00

10 100
Frequency (kHz)

Figure 4.30 Frequency dependence of core losses for various cut cores.

transverse field show a flat B-H loop and extremely low core loss below
several 10kHz. These nanocrystall ine Fe-Cu-Nb-Si-B toroidal cores are
particularly suitable for the inverter transformers because downsizing or high
efficiency can be realized due to large magnetic flux swing and very low core
loss (Ferch, 1998). Accordingly, these type cores are very suitable for
inverter transformers operated below several 10kHz.
4. 3. 4. 3 Active Filter Choke Coils. Smoothing Choke Coils
The good superimposed direct current characteristic is required for active filter
choke coils or smoothing choke coils. The low core loss in the high-frequency
range is also important to decrease the temperature rise of the choke coils for
switched-mode power suppl ies.
Figure 4. 31 shows the dc-superimposed characteristics of various cores.
The nanocrystalline soft ferromagnets are suited to choke coils for active filters
and smoothing choke coils due to high saturation magnetic flux density, and
low core loss. In addition, the noise arisen from magnetostriction can be
reduced as compared with Fe-based amorphous alloys because of low
magnetostriction of nanocrystalline soft ferromagnets. The net-shaped cores
with insulation layer consolidated from nanocrystalline flakes by hot-press
show good frequency dependence of lJi and two or three times the level of
saturation magnetization higher than Mn-Zn and Ni-Zn ferrites (Vincent and
Sangha, 1996). Hence, the dust cores fabricated from the nanocrystall ine soft
magnetic alloy powders or flakes are expected to be applied to active filter
choke coils or smoothing choke coils.
152 Yoshihito Yoshizawa

300
/=10 kHz
250
Nanocrystalline
Fe-Cu-Nb-Si-B(gap)
200
.A. Amorphous Fe base(gap)
Fe-AI-Si dust
:i: 150
Fe dust
~ Mn-Zn ferrite(gap)
100

o 2 468 10 12
Hdc(kA/m)

Figure 4.31 dc-superimposed characteristics of various cores.

4. 3. 5 Current Sensors

The small magnetic property change against temperature and high permeability
in the nanocrystalline soft magnetic materials are significant for current sensors
and electric leakage circuit breakers (Draxler and Styblikovo, 1996). If we
use the nanocrystalline cores of flat hysteresis curve with high saturation
magnetic flux density, high permeability and low remanence ratio for ground
fault interrupters, we can down-size the cores. Presently, the nanocrystalline
soft ferromagnets are replacing Permalloy as core materials for ground fault
interrupters due to their excellent properties.
Figure 4.32 shows the magnetization curves at 50 Hz and H m (maximum
magnetic field) dependence of P. at 50 Hz for the nanocrystalline FeCuNbSiB
alloys and Permalloy. The saturation magnetic flux density in the
nanocrystalline FeCuNbSiB alloy is higher than that of Permalloy. The applied
magnetic field characteristic in low-frequency range is comparable to that of
Permalloy. Moreover, they show good magnetic temperature stability.

4.3.6 Electromagnetic Shielding, Wave Absorbers

The nanocrystalline FeCuNbSiB alloys with high permeability, and high


saturation magnetic flux density are used as the magnetic shield materials. The
one that is combined with a resin is commercialized to make wide sheets, to
supplement brittleness of the ribbon after heat treatment. The sheet that is
combined with Cu and AI is developed to shield the electromagnetic wave.
Figure 4.33 shows the schematic drawing of a shield sheet cross-section
and applied field dependence of shield effects. The good shield effect is
Nanocrystalline Soft Magnetic Materials and Their Applications 153

No-field anneal
100 J=50 Hz

10-3 L.._ _---'-- L -_ _---'---_ _---'

10-2 10- 1 10 10' 102


H(A/m)

105

or
104

103
10-2 10- 1 10 10' 102
H(A/m)
Figure 4.32 Magnetization curves at 50 Hz and H m dependence of ~. at 50 Hz for
the nanocrystalline FeCuNbSiB alloys and Permalloy.

- - Polymer film (75 IlIn)


Adhesive layer (60 J.lm)
-----7"FINEMET" strips(20 mm)
/ (Nanocrystalline FeCuNbSiB alloy)

- - Polymer film (75 J.lm) or metal sheet (35 J.lm)

15
14 J=50 Hz
Nanocrystalline FeCuNbSiB alloy
12 ribbon sheet

~ 10
o
~ 8
"
"0

":.c: 6
C/J
4 Co-based amorphous alloy ribbon sheet
Co-based amorphous alloy flake sheet
2 ASTM Meas.:300 mm cube box
Ol.-_ _-..!.-_ _--L_ _- ' -_ _~-'--------'
0.4 2 4 10 20
Applied magnetic field (J.lT)

Figure 4.33 Schematic drawing of a shield sheet cross-section and applied field
dependence of shield effects.
154 Yoshihito Yoshizawa

obtained by using the nanocrystalline FeCuNbSiB alloy sheet, because of its


high saturation magnetic induction and high permeability.
The wave absorber which is combined the nanocrystalline Fe-Cu-Nb-Si-B
alloy powder with a resin, exhibits good properties in the frequency range of
GHz band.
Figure 4.34 shows the frequency dependence of complex permeability for
various powder-polymer composite sheets for wave absorbers. The
nanocrystalline FeCuNbSiB powder-polymer composite sheet exhibits high
complex permeability in the high-frequency range. Therefore, the
nanocrystalline FeCuNbSiB powder-polymer composite sheet shows high return
loss, as shown Fig.4.35.

Flaky nanocrystalline FeCuNbSiB powder+Polymer


S parameter method
Spherical Ni-Zn ferrite powder+Polymer

10'----------"..,,--------""-,,--------'-'--'
10' 102 103 104
Frequency (MHz)

Figure 4.34 Frequency dependence of complex permeability for various powder-polymer


composite sheets for wave absorbers.

-5

~
~
~
(Flaky FeAISi powder)
y
] -10 Conventional alloy powder-polymer
composite sheet, 2 mm

-15 Nanocrystalline alloy powder-polymer


composite sheet, 2 mm
(Flaky nanocrystalline FeCuNbSiB powder)
-20'--_---'-_ _---'- '--_----'-_ _-'-_ _---'
0.1 0.2 0.4 1 2 4 10
Frequency (GHz)

Figure 4. 35Frequency dependence of return loss for nanocrystalline FeCuNbSiB powder-


polymer composite sheet and FeAISi powder-polymer composite sheet.
Nanocrystalline Soft Magnetic Materials and Their Applications 155

4. 4 Conclusions

The nanocrystalline soft magnetic materials exhibit attractive magnetic


properties and a unique microstructure, which cannot be obtained in
conventional soft magnetic materials. Although many studies have been carried
out after the discovery of nanocrystalline Fe-based soft magnetic materials,
the relationship between soft magnetic properties and nano-structure is not
fully understood. Accordingly, research and development on magnetic
properties and nanostructure will be continued for nanocrystalline soft magnetic
materials from now on. Finally, nanocrystalline soft magnetic materials are
expected to progress as the materials for high-performance magnetic parts,
which matches the trend of increasing operating frequency in electronic
devices, noise reducing, down sizing, and energy saving.

References
Alben, R. , J. J. Becker and M. C. Chi. J. Appl. Phys. 49: 1653(1978)
AIIia, P. , C. Beatrice and F. Vinai. Appl. Phys. Lett. 59: 2454(1991)
Brovko, I., P. Petrovic, M. Zatroch and M. Konc. Key Engineering
Materials 81-83: 183(1993)
Duhaj, P. , P. ~vec, D. Janickovic and I. Matko. Mater. Sci. Eng. A 133:
398(1991)
Draxler, K. and R. Stybllkova. J. Magn. Magn. Mater. 157/158: 447(1996)
Ferch, M. PCIM '98 Japan Proc. 259(1998)
Fujii, Y., H. Fujita, A. SekiandT. Tomida. J. Appl. Phys. 70: 6241(1991)
Fukunaga H. , T. Eguchi, K. Koga, Y. Ohta and H. Kakehashi. IEEE Trans.
Magn. 26: 2008( 1990)
Hasegawa, N. and M. Saito. J. Jpn. Inst. Met. 54: 1270(1990a)
Hasegawa, N. and M. Saito. J. Mag. Soc. Jpn. 14: 313(1990b)
Hasegawa, N. and M. Saito. J. Magn. Magn. Mat. 103: 274(1992)
Herzer, G. IEEE Trans. Magn. 25: 3327(1989)
Herzer, G. IEEE Trans. Magn. 26: 1397(1990)
Herzer, G. Proc. Int. Symp. On 3 -0 Transition-Semi Metal Thin Films,
Sendai, 130(1991)
Herzer, G. J. Magn. Magn. Mat. 112: 258(1992)
Herzer, G. J. Mater. Eng. Performance 2: 193(1993)
Herzer, G. Mater. Sci. Eng. A 181/ A 182: 876 ( 1994a)
Herzer, G. IEEE Trans. Magn. 30: 4800( 1994b)
156 Yoshihito Yoshizawa

Herzer, G. Scr. Metall. Mater. 33: 1741(1995)


Herzer, G. J. Magn. Magn. Mater. 157/158: 133(1996)
Hoffmann, H. J. Appl. Phys. 35: 1790 ( 1964)
Hoffmann, B. and H. KronmUlier. NanoStructured Materials 6: 961(1995)
Hono, K. , A. Inoue and T. Sakurai. Appl. Phys. Lett. 58: 2180 ( 1991)
Hono, K., K. Hiraga, Q. Wang, A. Inoue and T. Sakurai. Acta metall.
mater. 40: 2137(1992)
Hono, K., J-L Li, Y. Ueki, A. Inoue and T. Sakurai. Appl. Surf. Sci. 67:
398(1993)
Hono, K. , D. H. Ping, M. Ohnuma and H. Onodera. Acta mater. 47: 997
(999)
Kataoka, N. , T. Matsunaga, A. Inoue and T. Masumoto. Mater. Trans. JIM
30: 947(1989)
Koba K., D. Arakawa, M. Fujieda, K. Ikegami, Y. Ishi, K. Kanai, C.
Kubota, S. Machida, Y. Mori, C. Ohmori, K. Shinto, S. Shibuya, A.
Takagi, T. Toyama, T. Uesugi, T. Watanabe, M. Yamamoto, T. Yokoi
and M. Yoshii. Review of Scientific Instruments 70: 2988(1999)
Komoto, O. , N. Uchida, E. Aoyagi, T. Choh and K. Hiraga. Mater. Trans.
JIM31: 820(1990)
Kraus, L. , K. Zaveta, O. Heczko, P. Duhaj, G. Vlasak and Schneider. J.
Magn. Magn. Mat. 112: 275(1992)
Mori, Y. , M. Fujieda, K. Koba, H. Nakayama, C. Ohmori, K. Saito, Y.
Satoh, Y. Tanabe, A. Takagi, Y. Toda, T. Uesugi, M. Yamamoto, T.
Van and Y. Yoshii. Proc. of European Accelerator Conference.
Stockholm, 299(1998)
MUlier, M., N. Mattern, L. IIlgen, H. R. Hilzinger and G. Herzer. Key
Engineering Materials 81 - 83: 221( 1993)
MUlier, M. and N. Mattern. J. Magn. Magn. Mater. 136: 79(1994)
MUlier, M. , N. Mattern and U. KUhn. J. Magn. Magn. Mater. 157/158: 209
(1996)
Nakajima, S. , S. Arakawa, Y. Yamashita and M. Shiho. Nuclear Instruments
and Methods in Physics Research A 331: 318(1993)
Nakajima, S. , H. Miki, S. Kubota and M. Sakaguchi. Hitachi Metals Tech.
Report. 15: 37 ( 1999)
Nakanishi, K., O. Shimizu and S. Yoshida. J. Mag. Soc. Jpn. 15: 371
( 1991)
Neel, L. J. Phys. Radium 15: 225 ( 1954)
Reininger, T. , B. Hofmann and H. KronmUlier. J. Magn. Magn. Mater. 111:
L220(1992)
Saito, K., J. I. Hirota and F. Noda. Nuclear Instruments and Methods in
Physics Research A 402: 1(1998)
Sajiki, K., T. Nisizaka, S. Nakajima and S. Watanabe. IEEE J. Quantum
Nanocrystalline Soft Magnetic Materials and Their Applications 157

Electron. 31: 2183(1995)


Sakurai, M., M. Matsuura, S. H. Kim, Y. Yoshizawa, K. Yamauchi and K.
Suzuki. Mater. Sci. Eng. A 179/180: 469(1994)
Sawa, T. and Y. Takahashi. J. Appl. Phys. 67: 5565(1990)
Schafer, R. , A. Hubert and G. Herzer. J. Appl. Phys. 69: 5325 ( 1991)
Suzuki, K., N. Kataoka, A. Inoue, A. Makino and T. Masumoto. Mater.
Trans. JIM 31: 743( 1990)
Suzuki, K. , G. Herzer and J. M. Cadogan. J. Magn. Magn. Mater. 177-
181: 949( 1998)
Suzuki, K. and J. M. Cadogan. Physical Review B 58: 2730 (1998b)
Suzuki, K., A. Makino, A. InoueandT. Masumoto. Jpn. J. Appl. Phys. 30:
Ll729(199l)
Taneko, N. , Y. Shimada, K. Fukamichi and C. Miyakawa. Jpn. J. Appl.
Phys. 30: L195(199l)
Tomida, T. Mater. Sci. Eng. A 179/180: 521(1994)
Trasov, L. P. Phys. Rev. 56: 1231(1939)
Ueda, Y., S. Ikeda and K. Minami. Mater. Sci. Eng. A 181/182: 992
(1994)
Vincent, J. H. and S. P. S. Sangha. GEC Journal of Research 13: 2(1996)
Watanabe, H., J. Saito and M. Takahashi. J. Mag. Soc. Jpn. 17: 191
(1993)
Watanabe, K. , M. Mizuno, S. Nakajima, T. Iimura and Y. Miyai. Review of
Scientific Instruments 69: 4136( 1998)
Yoshizawa, Y., S. Oguma and K. Yamauchi. J. Appl. Phys. 64: 6044
(1988a)
Yoshizawa, Y., K. Yamauchi and S. Oguma. European Patent Application
0,271,657 (1988b)
Yoshizawa, Y. and K. Yamauchi. J. Japan Inst. Met. 53: 241( 1989a)
Yoshizawa, Y. and K. Yamauchi. J. Mag. Soc. Jpn. 13: 231(1989b)
Yoshizawa, Y. and K. Yamauchi. IEEE Trans. Magn. 25: 3324(1989c)
Yoshizawa, Y. and K. Yamauchi. Mater. Trans., JIM 31: 307(1990a)
Yoshizawa, Y. and K. Yamauchi. J. Mag. Soc. Jpn. 14: 193(1990b)
Yoshizawa, Y. and K. Yamauchi. Mater. Sci. Eng. A 133: 176(1991a)
Yoshizawa, Y. and K. Yamauchi. Mat. Res. Soc. Symp. Proc. 232: 183
(1991b)
Yoshizawa, Y. and K. Yamauchi. J. Japan Inst. Metals 55: 588(1991c)
Yoshizawa, Y. and K. Yamauchi. Mater. Sci. Eng. A 133: 176 ( 1991 d)
Yoshizawa, Y. and K. Yamauchi. In: Current Topics in Amorphous Materials:
Physics and Technology. eds. Elsevier Science Publishers B. V., p.234
(1993)
Yoshizawa, Y. , H. Mori, S. Arakawa and K. Yamauchi. J. Mag. Soc. Jpn.
19: 457(1995)
158 Yoshihito Yoshizawa

Yoshizawa, Y., H. Miki, T. Meguro, Y. Bizen, S. Nakajima and S.


Arakawa. Hitachi Metals Tech. Repo. 13: 25(1997)
Yoshizawa, Y. Mater. Sci. Forum 307: 51C 1999)
Zaveta, K. , K. Jurek, J. Schneider, R. Hesske and L. IIlgen. Mater. Sci.
Forum 62-64: 525( 1990)
Zhang, Y. , K. Hono, A. Inoue, A. Makino and T. Sakurai. Acta mater. 44:
1497(1996)
5 Spin-Density Waves and Charge-Density Waves
in Cr Alloys

A. J. A. de Oliveira and P. C. de Camargo

5. 1 Introduction

Chromium metal is an itinerant antiferromagnet, which orders as a spin-density


waves (SOW) incommensurate with the lattice. The magnetic phase diagrams
present three distinct phases: The paramagnetic phase (P) above Neel
temperature TN; the SOW phase with transversal polarization (AF 1 ) , between
TN (311 K) and spin-flip temperature T SF' which for Cr is 123. 5 K; and the
longitudinal polarization (AF2 ) , below T SF ' In the AF 1 phase the wave vector
Q is perpendicular to spin polarization vector (S), while in the AF 2 phase, the
polarization is parallel to Q. Therefore, the AF 1 phase is characterized by
magnetic domains of the type (QI ,Sj)' with i, j (i, j = H, K, L-directions of
reciprocal space), and orthorhombic symmetry. The AF 2 phase has the
magnetic domains (0 1 , Sj) with i = j, and, therefore shows tetragonal
symmetry (Fawcett, 1988; Fawcett et al. , 1994).
The first suggestion that chromium is an antiferromagnet was proposed by
Neel (1936), to explain the change in electrical resistivity as a function of the
pressure observed by Brigdman (1932). Only 20 years latter, in 1953, its
itinerant and incommensurate nature was discovered by Shull and Wilkinson
(1953), through elastic neutron diffraction experiments. Since then, a variety
of the experimental data on Cr and its alloys has been accumulated. The SOW
in Cr is regarded as a classical and canonical system to study incommensurate
states and provides a testing ground to check several theoretical proposals.
The itinerant nature of all antiferromagnetic SOW alloys brings up a wide
range of possibilities to explore the relation between the Fermi level position
and the density of states (Fishman, 2001). The electrons/atom ratio can be
changed by alloying with other transition metals, showing quite dramatic
effects on magnetic phase transition temperatures that can reach hundreds of
kelvins, with only one percent of alloying. External pressure, internal
stresses, and magnetic field also significantly alter the transport properties,
elastic constants, and magnetic response.
The phenomena of spin-density waves is present in many different
systems, because the SOW are broken-symmetry ground states of metals
160 A. J. A. de Oliveira and P. C. de Camargo

present, for example, in several organic linear-chain compounds, as


(TMTSF)2PF6(Gruner, 1994). The development of the SOW state, in these
systems, opens up a gap in the single-particle excitation spectrum, and the
ground state is close to that of an antiferromagnet, as shown by a wide range
of magnetic studies. The similarity of the ground state with other systems,
which present superconductivity and charge-density waves, stimulates the
comprehension of this peculiar magnetic state. Nevertheless, chromium is the
best prototype for understanding spin-density waves systems, because its
magnetism comes from spin contribution only.
Another motivation that encourages theorists and experimentalists to
understand the nature of this peculiar type of the magnetism is the possibility of
its connection with superconductivity (Gabovich et al. , 2001). Mason et al.
( 1992) observed incommensurate SOW fluctuations in the high-temperature
superconductor of La2- x Sr x CU04 very similar to those seen in the paramagnetic
phase of chromium (Noakes et aI., 1990) and Cr-O. 2 at. % V (Noakes et al. ,
1990). More recently, in the superconductors of the type borocarbides (Lynn et al. ,
1997), the coexistence of superconductivity and incommensurate SOW, below and
above the critical temperature was observed. In fact, many years ago, Overhauser
( 1962) suggested the nesting electron and hole Fermi surfaces, and the first
conclusion from the similarities in the dynamics susceptibility was that the nesting of
Fermi surfaces can be associated with the superconductivity, which occurs also in
the cuprate' s superconductor. These aspects are discussed in detail by Anderson
(1997), where the author proposes a connection of spin fluctuations and Fermi liquid
theory as an explanation of the behavior observed in spin-density waves and
superconductor phase.
The persistence of the existence of the incommensurate SOW in magnetic
multilayers , as Fe/Cr, is another interesting aspect concerning Cr and
antiferromagnetic alloys. Since the discovery of giant magnetoresistance in
multilayers of Fe/Cr by Baibich et al. (1988), great effort has been done in
understanding the magnetic coupling through chromium antiferromagnetic
layers (Zabel, 1999).
Finally, the Neel transition in Cr is often referred to as weak first order.
When observed through certain properties as in magnetization measurements,
it resembles a second order transition, even though neutron scattering (Grier
et aI., 1985) and thermal expansion (Fawcett et aI., 1986) experiments
confirm its first order nature. In fact, the small, however clear, value of the
latent heat involved, of about 1.2 J/mol (Williams et aI., 1979), is the
reason for this unusual classification.
Approximately 70 years after the discovery the antiferromagnetism in Cr,
much progress has been made in understanding the mechanisms involved;
however, new problems and new interconnections have been established. In
fact, the magnetism of chromium, with its SOW still challenges both
theoreticians and experimentalist of the 21 st century.
Spin.Density Waves and Charge-Density Waves in Cr Alloys 161

S. 2 Magnetic Properties of Antiferromagnetic Cr and Cr Alloys

5.2.1 Magnetic Phase Diagrams


Antiferromagnetism in Cr is produced by the Coulomb attraction between
electrons and holes on nested portions of the Fermi surface. A two-dimensional
projection of the Fermi surface is shown in Fig. 5. 1. The electron jack
centered at the position in reciprocal space r point and the hole octahedra
centered at the H points are roughly nested by three sets of nesting wave
vectors, of which it draws the set Q along the z-axis. The selected Hand r

points are joined by half of reciprocal-lattice vector G z = (4;)z, where a is


the lattice parameter for the conventional cubic unit cell. Because the hole
octahedron is slightly larger than the electron jack, the nesting wave vectors
can be written as Q = (2;) (1 0), where the Fermi surface mismatches 0

is about O. 05 for pure Cr. The 0 increases with the introduction of the
transition metals at the left of chromium in the periodic table, as for V and Ti,
for example. On the other hand, 0 decreases and vanishes for certain
concentration of metals, as with the introduction of 3d-metal impurities like
Mn, Fe, Re, changing the SDW to commensurate, i. e., 0 = O. More details
on the change of magnetic behavior shall be discussed in Section 2.2.

Figure 5. 1 A two-dimensional projection of the Fermi surface of Cr, showing the electron
jack centred at the r point, the hole octahedron cenetrd at the H points, hole ellipses at
the N points and one set Q of nesting wavectors (Fishman, 2001).
162 A. J. A. de Oliveira and P. C. de Camargo

Chromium alloys exhibit many magnetic phases, as the commensurate


antiferromagnetic (AFo ), in alloys of Cr-Mn (Geerken et ai., 1982), Cr-Si
(Benediktsson et ai., 1982; Endoh et ai., 1982), Cr-Fe (Benediktsson
et al., 1982), Cr-AI (Benediktsson et al., 1975), Cr-Co (Benediktsson
et ai., 1975), Cr-Ru (Sidek et ai., 1993), ferromagnetic in Cr-Fe (Burke
and Rainford, 1983, 1978), and superconductor in Cr-Re (Muheim and MOiler,
1964), Cr-Ru (Nishihara et ai., 1986).
Antiferromagnetic state in Cr alloys may present three distinct ordered
phases: transverse SOW (AF 1)' longitudinal SOW (AF2 ), and commensurate
SOW (AFo ). Above the Neel temperature all Cr alloys are paramagnetic (P).
The neighborhood of triple point, AF o , and AF 1 phases are of great physical
interest, and there are many microscopic theories to explain this point in the
magnetic phase diagram. On the other hand, the spin-flip transition does not
have a microscopic theory for the understanding of this magnetic phase
(Fawcett et al. , 1994). The magnetic phase diagrams for Cr alloys have been
studied as functions of concentration, pressure, temperature, and magnetic
field. Anomalies in different physical properties characterize the Neel
transition. Parameters like thermal expansion and neutron intensities, are
frequently observed as continuous and slow change quantities at the Neel
temperature, where they are expected to be discontinuous. This is likely due
to internal defects or chemical inhomogeneities that broaden the weak first
order transition.
Lomer ( 1962) recognized that the large amplitude of the SOW is
connected with peculiar geometric features of the Fermi surface of Cr, which
permit nesting between electron and holes sheets having similar shape,
namely, the electron and hole octahedra situated at the rand H points,
respectively in the Brillouin zone of a bcc crystal of Cr. The hole octahedron is
sl ightly larger than the electron one. They nest each other effectively to induce
SOW instability. When transition metal impurities are introduced, rigid band
model explains the changes in SOW state of Cr. For example, the introduction
of V, which has a smaller electron/atom ratio than that of Cr, increases the
size of hole octahedron and decreases that of electrons. As a consequence,
the incommensurate parameter {j increases and the Neel temperature and
ampl itude of SOW decreases. For concentration of 3. 8 at. % V in Cr, the
mismatch {j is large enough to make the SOW unstable even at 0 K, and the
Neel transition vanishes. On the other hand, when introducing Mn, which has
an electron/atom ratio larger than that of Cr, the opposite occurs, i. e. , the
electron octahedron increases and the mismatch {j decreases, such that the
SOW becomes stable at higher temperatures. For 0.3 at. % Mn, the Neel
transition is at TN = 325 K (the triple point in phase magnetic diagram), {j
vanishes, and the SOW state becomes commensurate.
The magnetic behavior around the triple point of Cr-Mn was explained by
Machida and Fuj ita (1984) using a two-band imperfect-nesting model within a
mean-field approximation. This result successfully explains various
Spin-Density Waves and Charge-Density Waves in Cr Alloys 163

experimental data on Cr and Cr-rich Cr-Mn and Cr-V alloys.


The Neel temperature dependence on impurities concentrations has been
observed through different physical properties. In particular, the number of
conduction electrons is reduced due to the gap formed in the antiferromagnetic
phase, and as a consequence, the resistivity rapidly increases as the AF 1
state is formed, as shown in Fig. 5. 2 .
68

67

66
/
/
65
/
Cl.. /
64 /
/
/ Hc =55 kG
63

Transverse
62

61
260 270 280 290 300 310 320 330
T(K)

Figure 5. 2 The electrical resistance of a single crystal of Or cooled through the Neel
temperature in a magnetic field. The current and field were both in (100) directions
(Trego and Mackintosh, 1968).

For the cases of transitions between AF 1 and AF o phases (Tic-transition


temperature of incommensurate and commensurate phase) and of the spin-flip
temperature (T SF)' the spin configuration changes, but the antiferromagnetic
state is maintained and the electric resistivity no longer detects the transition
temperature. TIc has often been determined from anomalies in the temperature
dependence of the magnetoelastic properties, namely, thermal expansion,
sound velocities, and magnetic measurements. The most direct measurement
is definitively the neutron's diffraction that shows the onset of AF 1 phase
through the appearance of incommensurate Bragg peaks.
The identification of the correct temperature of a phase transition, from an
anomaly of a physical property involves a problem of defining a clear criteria.

*'
In the case of the temperature dependence of the electrical resistivity, p, it
may show a true minimum, but the Neel temperature is best determined as the

minimum of which corresponds to the inflexion point as indicated in


164 A. J. A. de Oliveira and P. C. de Camargo

Fig. 5.2. This procedure was proposed by Arajs et al. (1980), who compared
values of TN in a Cr-Re alloy system with values obtained by neutron diffraction
(Lebech and Mikke, 1972).
The Neel transition, in Cr-V alloys, was identified using ac magnetic
susceptibility measurements by de Oliveira et al. (1996a, 1996b). With this
technique it has been possible to simultaneously obtain the real component of
magnetic susceptibility (X'), that is proportional to magnetic moment of the
sample, and the imaginary component (X") associated to the energy
dissipation. For insulators (X") is proportional to the energy dissipated due to
the re-arrangement of magnetic moments, but in metallic samples it may be
considered inversely proportional to electrical resistivity (de Faria et al. ,
1995). In this case, the Neel transition can also be defined as the inflection
a"
point for L. Similarly, other simultaneous measurements show that the Neel
aT
point is properly defined in dc magnetic susceptibility, Xdc = M/H, on the

inflection of aa\dc point, as shown in Fig. 5.3.


p

3.30
~

?'" 3.25
~
8
",'" 3.20
I
o

o 50 100 150 200 250 300 350 400


Temperature (K)
Figure 5.3 Low magnetic field susceptibility as a function of temperature for a Cr single
crystal. H = 500 Oe applied parallel to (100) direction. The arrows denote the Neel
temperature (TN) and spin-flip temperature (T sF ).

The magnetic phase transitions in chromium alloys are strongly dependent


on sample quality. Chemical impurities may cause changes in the expected
transition temperature. Internal stresses and inhomogeneities would cause a
broadening effect.

5. 2. 2 Magnetic Phase Transition


5.2.2. 1 Neel Transition
The nature of the Neel transition in pure Cr is not completely clear. The fact
Spin-Density Waves and Charge-Density Waves in Cr Alloys 165

that the Neel transition is of weak first order has been discussed by several
different approaches, but the question of its origin has not been resolved. The
observation of an abrupt decrease in the intensity, the elastic neutron
scattering, and the discontinuity in the thermal expansion are characteristic of
the first order transition. But this transition is called weak and intervenes in
what appears to be the approach to a continuous transition, as the order
parameter decreases with increasing temperature. On the other hand, the
measurements of the specific heat exhibit a clear small latent heat of 1.2 J/mol
at the Neel temperature (Fawcett, 1988).
It is worth mentioning that, as far as the authors know, there is no good
reference available for the temperature dependence of magnetic susceptibility
of chromium. The curve presented by (Fawcett, 1988; Bender and Muller,
1970) shows the expected peak at the Neel temperature; however, it has not
been observed by other authors. There are difficulties in establishing a good
reference for magnetic susceptibility as it is discussed in Section 5. 3. 1. A
typical case of introducing transition metals in chromium in order to understand
its physical properties is the introduction of V. Vanadium is an ideal doping
because of its similarity with paramagnetic Cr and because it causes a linear

dependence of Neel temperature such that aa:


N
= - 80 K/at. %, while the
introduction of Nb, which is a 4d metal or Ta (5d metal), both with the same
electronic structure as V, cause a aa:N
= - 70 K/at. % and - 100 K/at. %
respectively. The addition of Ti which has two electronslatom less than Cr
shows a aa: = -
N
160 KI at. %. Therefore, the Neel temperature dependence
on the electronl atom ratio is consistent with the rigid band picture.
On the other hand, the introduction of elements of the right side of Cr in
the periodic table increases the Neel temperature. The typical example is the
introduction of Mn or Re in Cr, where aa: =50 and 80 K/at. %, respectively,
N

where the relationship electron/atom is higher than one. The exceptions are
Fe and Co that decrease the Neel temperature at a rate of - 25 and
20 KI at. %, respectively. This behavior can be explained by the fact that Fe
and Co have an intrinsic magnetic moment that interacts with the SOW
decreasing its amplitude (Fawcett et al. , 1994).
There is renewed interest in the nature of the Neel transition in Cr and its
dilute alloys. In pure Cr the transition is weakly first order, with an abrupt
onset of the SOW at the Neel temperature, as seen in neutron diffraction
(Noakes et aI., 1990). Similar behavior was seen in the thermodynamic
properties, specific heat (Williams et aI., 1979) and thermal expansion
(Fawcett et al. , 1986), with the picture for the elastic moduli being obscured
by the strong attenuation close to TN for those that show large effects
associated with the SOW (Muir et aI., 1987b).
166 A. J. A. de Oliveira and P. C. de Camargo

Critical fluctuations are seen in pure Cr around the Neel transition in all
these properties, but in the ordered phase below TN the rapid increase in the
magnitude of the order parameter with decreasing temperature soon dominates
their temperature dependence. There is a little evidence for critical fluctuation
in the transport properties, but below TN the rapid variation of the resistivity
and the thermoeletric power in their ordered phase are associated with the
growth of the SOW with decreasing temperature. The anisotropy of the
resistivity in the ordered phase shows that its rapid increase below TN is due
to the opening of a gap at the Fermi surface due to the formation of electron-
hole pairs, in accordance with Overhauser' s analysis (Overhauser, 1962) of
the origin of the SOW in Cr. In summary, in pure Cr there is a jump in the
amplitude of the SOW at the Neel temperature, which produces a first-order
change in the entropy (Williams et ai., 1979) and the strain (Fawcett et al. ,
1986).
In the case of Cr-V alloys, Fawcett et al. (1986) and de Camargo et al.
(1988) have observed that the alloy Cr-O. 2 at. % V exhibits a second order
transition in the Neel temperature, as shown by thermal expansion and neutron
scattering. The nature of the second order transition in Cr-V alloys is not
totally understood, but some insight is gained following Fishman and Liu
( 1992 ), who proposed that the introduction of V impurities continuously
decreases the amplitude of the SOW, and as a consequence, so does that of
the charge-density waves(COW). This effect changes the weak first order to a
second order transition. On the another hand, Cr-O. 18 at. %Re alloy the Neel
transition remains first order (Fawcett and Noakes, 1993), but this is believed
to be due to the close proximity of the system to the phase boundary between
the incommensurate and commensurate SOW phase.
The magnetoelastic interactions in chromium are such that the Neel
transition does not show dependence on magnetic field. Even for fields up to
20 T, no effect has been observed in the Neel temperature. On the other hand,
there is a strong pressure dependence of the Neel temperature, about
-0.5K/GPa (Mitsui and Tomizuka, 1965). For some Cr alloys, the
dependence of the Neel temperature with pressure is even stronger as in Cr-
1. 85 at. % Si, where aa~N is approximately -500 K/GPa.
The Neel temperature of nanocrystalline chromium exhibits important
changes. Fitzsimmons et al. (1993, 1994) identified two instances where
nanocrystalline chromium samples were not antifferomagnetically ordered even
at temperatures as low 20 K, indicating that the Neel temperatures of these
samples are suppressed drastically when compared with a single crystal
chromium.
5.2.2.2 Spin-Flip Transition
At temperature of T SF = 123 K, chromium exhibits a first-order phase
transition, where the polarization direction in the transverse SOW phase (AF 1 )
Spin-Density Waves and Charge-Density Waves in Cr Alloys 167

"flips" into the direction of the wave vector Q in the longitudinal SOW phase
(AF 2 ) This transition is clearly observed in neutron-diffraction measurements,
because of the dependence of scattering intensity on the angle between the
polarization direction and momentum transfer to the scattered neutrons. This
transition is also clearly observed by X-ray magnetic scattering measurements,
when the intensity of the magnetic peak vanishes to zero for T < T SF when the
spin becomes parallel to Q (Hill et ai., 1995).
The AF 2 phase has been poorly investigated in Cr alloys. So far, there is
no microscopic theory to adequately treat the spin-flip transition, though it can
be modeled by use of a Landau-type free energy, as shown by Allen and Young
(1975, 1977) in their study of the magnetic anisotropy effects on elastic
properties.
The spin polarization vector (S), but not the periodicity wave vector
(Q), changes direction at T SF' suggesting that in the single-Q state there is a
uniaxial magnetic anisotropy with respect to the Q direction, which lies along
one of the cubic axes of the paramagnetic state. This uniaxial anisotropy is in
addition to the crystalline axes, analogous to the cubic anisotropy found in
ferromagnets having cubic symmetry in the paramagnetic state (Allen and
Young, 1977).
The spin-flip transition has been observed in several Cr alloys systems,
including Cr-Ti (Chiu et ai., 1971), Cr-W (Hedman et ai., 1974), Cr-AI
(Alberts and Lourens., 1984; Baran et ai., 1992), Cr-Ru (Alberts et al. ,
1992), Cr-Pt (Alberts and Lourens, 1988), Cr-Mn (Koehler et al. , 1966),
Cr-Re (Lebech and Mikke, 1972), Cr-Si (Endoh et al. , 1982), Cr-Be (Iida
et al. , 1980, Cr-Ge (lida et ai., 1980, and Cr-V (de Camargo et al. ,
1990; de Oliveira et ai., 1995). In all these alloys, the spin-flip temperature
decreases with respect to pure Cr. The Cr-alloys, which exhibit commensurate
phase, do not present spin-flip transition, because of the cubic symmetry,
while the AF 1 phase presents orthorhombic symmetry in the case of
incommensurate phase.
In contrast with the Neel temperature, that is insensitive to magnetic fields
up to 20 T, the spin-flip temperature strongly depends on the applied magnetic
field. Steinitz (1969), through thermal expansion measurements, have
determined the effect of single-Q state on spin-flip transition. The lattice
parameter shows a strong jump at T SF when the sample is cooled after the
preparation of the single-Q state. The same effect is not observed for the
sample cooled without applied magnetic field. Preference of Q to be oriented
on a particular direction is given by any anisotropy caused for example, by
external magnetic fields, pressure or even sample geometry (Fawcett, 1988;
Hill et al. , 1995). Street et al. (1968) showed that a field-cooled sample
remains single-Q, and the distribution of Q-domains in a poly-Q sample remains
unchanged even for temperature excursions taking the sample back and forth
through the AF 1 and AF 2 phases. Also, according to those authors, cooling
through the Neel transition with a tensile stress along [1 0 0] results in a single-
168 A. J. A. de Oliveira and P. C. de Camargo

Q domain in this direction. Recently, de Oliveira et al. (2001) have observed


that for low magnetic field near T SF' the values of magnetic susceptibility as a
function of temperature in the AF 2 phase are larger than those in the AF 1 phase.
On the other hand, for larger magnetic fields the magnetic susceptibility in AF,
phase is larger than in the AF 2 phase. This behavior can be better understood if
one considers that for magnetic fields above 6 kOe, where S-switching starts to
occur with saturation happening above 30 kOe for T> TsF(de Oliveira et al. ,
2001). Present results clearly indicate that a complete investigation of the
magnetic properties of chromium considering magnetic field effects and sample
quality is still lacking. In fact, recent results also bring up the question of how
surface preparation or sample external geometry may affect the domains
configuration and size.

5.3 Local Moments in Cr Alloys

5.3.1 Magnetic Susceptibility


Chromium, as a transition metal, shows paramagnetic behavior for high
temperatures, and exhibits the characteristic behavior of the many-body
system. The magnetic susceptibility in paramagnetic phase is of the Pauli
susceptibility type, with small variation with the temperature. Because of the
itinerant nature of the magnetism of pure Cr, its sub-lattice magnetic moment,
around 0.43 JJB' does not give rise to a Curie-Weiss behavior above TN' in
opposition to the 3d ferromagnetic metals such as Fe, Ni and Co, that show
this behavior above the critical temperature.
Magnetic susceptib iIity , X do ' is one of the most important physical
properties for the characterization of magnetic materials; however, this
aspect is poorly explored in Cr and its alloys. In fact, as far as magnetic
susceptibility of Cr and of its alloys is concerned, many contradictions are
observed in literatures. Because of its very small absolute value at room
temperature, which is of the order of 3.26 x 10- 6 emu/(g Oe) (or 2.55 x
10- 4 m3 /kg), and its very small temperature dependence, only recently with
SQUID magnetometers was it possible to perform a reliable magnetic
characterization of Cr. In addition to the small signal there is a departure from
the expected result for a classical antiferromagnet, which is a peak at the Neel
temperature. In fact, Bender and Muller (1970) have found this expected peak
at TN' but no one could confirm their result.
The X de and the X 80 magnetic susceptibilities may give raise to different
information, so, it is worthwhile to consider these quantities carefully. The
quantities X de defined as the ratio between magnetization M and a fixed
Spin-Density Waves and Charge-Density Waves in Cr Alloys 169

appl ied magnetic field H, and X ae = (~~ ) H,T


where h is the ac magnetic field of
excitation, could only coincide if M was a linear function of Hand frequency-
independent. In addition, the differential susceptibility (Xae)' when contrasted
to the dc measurements (X de)' provides richer information on and a better
definition of the magnetic susceptibility around the phase transition. Figure 5.3
shows the dc magnetic susceptibility as a function of temperature for a highly
pure and good crystalline quality sample (FWMH = 23"). For this sample we
have identified TN as the inflection point of the susceptibility as a function of
temperature. The spin-flip temperature is determined using the same criterion.
In 1966, Fedders and Martin (1966) proposed a model including the
"nesting" of spheres of electron and holes of the same radius kF(Fermi vector)
and same Fermi velocity VF that gives the essential features of the anisotropy
and temperature dependence of the magnetic susceptibility of an SOW system.
With this model, in the paramagnetic phase the magnetic susceptibility is
independent of the temperature and their value Xo is

(5.1)

Below TN the components of tensor susceptibility X.L' perpendicular to


polarization of SOW (S) decrease linearly with temperature,

(5.2)

Although, in the direction of polarization, the susceptibility is constant:

X//= XO' (5.3)

However, for polycrystalline samples of Cr-Ru, Arajs et al. (1975) observed


through magnetic investigation between the spin-flip temperature and TN that
the magnetic susceptibility present a quadratic temperature dependence:

X.L=X//=Xo 1+ [ ..r(kBT)2J
1'ik V (5.4)
3 F F

The introduction of the magnetic impurities in Cr results in the formation of


local magnetic moments, which in paramagnetic phases, obey the Curie-
Weiss law. Among all known Cr antiferromagnetic alloys, the Cr-Fe with low
concentration of Fe, is the only one for which the local moment persists into
the SOW phase. Friedel and Hedman (1978) have proposed an interpretation
of this particularity assuming that a single Fe atom is coupled to the SOW,
while Fe clusters interact so strongly with each other that they overcome the
coupling to the SOW.
On another hand, this behavior is not observed when Co and Ni are
introduced in Cr. Below TN the magnetic susceptibility of these alloys is
similar to that of pure Cr, apart from the change in the Neel temperature and
170 A. J. A. de Oliveira and P. C. de Camargo

the Curie-Weiss behavior observed in the paramagnetic phase.

5.3.2 Local Moments in Non-Magnetic Impurities-Local


Spin-Density Waves
The introduction of ferromagnetic elements in Cr may lead to a Curie-Weiss
behavior (CW) superimposed in antiferromagnetic and paramagnetic phases,
for example, in Cr-Fe (Suzuki, 1966; Hedgcock et a!., 1977) and Cr-Mn
(Aidun et a!., 1985) alloys. However, there is a possibility of formation of
local magnetic moments in Cr alloys with the introduction of non-magnetic
impurities, as predicted by Tugushev (1992). Tugushev' s model assumes the
establishment of local spin-density wave around lattice or chemical defects.
This local SOW may be stable, even at temperatures higher than TN. The
formation of local SOW around the defects creates a region of nonzero
magnetization M, where short-range magnetic order takes place. Taking F,
as the defect potential, N(O) as the density of states at the Fermi level, and
.1 t (0) as an order parameter proportional to the energy gap (.1), calculation
to first order in F 1 yields:
(5.5)
For the region where the local SOW decays exponentially away from the
defect, the temperature interval in which a local moment exists has been
estimated to be
.1 100 = T loe - TN ~ F~ (5.6)
where T loo is the temperature above which there would be no local SOW.
In the particular case of Cr-V alloys, this behavior has been observed by
Kondorskii et a!. (1981), Hill et a!. (1994) and de Oliveira et a!. (1996a,
1996b, 1997). In these alloys the Curie-Weiss behavior shows a strong
dependence applied magnetic field and on V concentration (CW disappears for
concentrations of V above O. 67 at. %) (de 01 iveira et ai., 1996b). For
concentration 0.2 at. %V and 0.4 at. %V the critical field (He) is of the order
10 kOe (de Oliveira et ai., 1997). For fields above this value, the magnetic
susceptibility of these alloys reproduces exactly that of pure Cr, as shown in
Fig. 5.3 .In Fig. 5.4 the Curie-Weiss behavior of Cr-x at. % V (x =0.1, 0.2,
o. 4) is shown to be suppressed for high magnetic fields.
Recently, de Oliveira et al. (2003) found that T loo = 340 and 370 K, for
Cr-O. 1 at. % V and Cr-O. 2 at. % V, respectively. These temperatures agree
with the model proposed by Tugushev (1992).
Using Eqs. (5.5) and (5.6) and experimental data for Cr-V alloys, one
can estimate the change in .1 100 due to the increase in V concentration. The
value of the energy gap in Cr-V and some other Cr alloys with the
incommensurate SOW structure, determined from the peak in the infrared
Spin-Density Waves and Charge-Density Waves in Cr Alloys 171

~ 1.75
<U
o
"0
E
~ 1.70
E
.,.
<U
I
o
2S 1.65
~
240 260 280 300 320 340
Temperature (K)
(a)
1.50

~ 1.48
"0
E
] 1.46
.,.<U
I
o 1.44
~
x
~
1 .....--
1.42

260 280 300 320


Temperature (K)
(b)
~

<i'
o 1.45
"0
E
"' 1.44
E
.,.<U
S 1.43
x
~

.g
~ 1.42

240 260 280 300


Temperature (K)
(c)

Figure 5.4 Magnetic susceptibility as a function of temperature for Cr-x at. % V. (a) x =
0.1, H=O.l kOe(1) , 1.0kOe(2), 2.0 kOe(3) , 20kOe(4). (b) x=0.2, H=2.0kOe
(1), 8.0kOe(2), 10.0kOe(3), 15kOe(4). (c) x=0.4, H=2.0kOe(1), 10.0kOe
(2), 12 kOe(3). The applied magnetic field suppressed the local spin-density waves
around Vimpurities (de Oliveira et al. ,1997).

reflectivity at low temperatures, has been shown Baker and Ditzenberger


( 1970) to be proportional to the Neel temperature:

..1 loc = 5.0k s TN (5.7)


172 A. J. A. de Oliveira and P. C. de Camargo

The change in de density of states at the Fermi level can be estimated


from the concentration dependence of the specific heat coefficient y, which
increases almost Iinearly with V concentration. So, .1 (x) in Cr-x at. % V
decreases and density of states, N (x), increases almost linearly with x. The
product of these value for x = O. 2 and O. 4, for example, stays almost
constant. According to Eq. (5. 5) this means that the value of the defect
potential F 1 has to be proportional to the magnetic moment. For the case of
Cr-V alloys with concentrations 0.2 at. % V and 0.4, in low fields, they carry
magnetic moments of - 1. 5PB and -1. OPB' respectively (Hill et a!. , 1994).
Therefore, according to Eq. (5.6), the value of .1 loc for the first alloy should
be more than twice as large as that for the second one. So the temperature
interval between T ,oc and TN in Cr-V alloys seems to decrease rapidly with
increasing V concentration. It is important that, according to the three-
dimensional Tugushev' s model (1992), some minimal value of the defect
potential F 1m1n exists, which limits the possibility of the local transition being
realized. Therefore, the local magnetic moments in Cr-V systems disappear
for x~0.67.
One might say that the magnetic field would cause a decrease in T loc ,
which would then collapse onto TN. de Oliveira et a!. (2003) observed that
T ,oc presents the linear dependence of magnetic field. The magnetic moment
associated to Curie-Weiss behavior decreases exponentially as a function of
magnetic field, indicating that the magnetic field may inhibit the formation of
local moments. For lower fields the jump in magnetic susceptibility, as shown
in Fig. 5.4, is pronounced, i. e, there is a higher density of local moments (or
local SDW) around the V atoms. As the field is enhanced, the jump decreases
as a consequence of lower moment densities.
Nevertheless, the appearance of local magnetic moments depends both
on the V concentration and on the magnetic field. Therefore, one can sketch a
phase diagram where a critical line H L separates the CW phase from
electronic paramagnetism. Figure 5. 5 is a proposition of such a phase
diagram. de Oliveira et a!. (1996b) observed that local magnetic moment
does not manifest for x ~ 0 . 67, even for very low fields (7 Oe), suggesting
that Hc(x) vanishes within the interval 0.4~x~0.67.
Besides, one can recall the theory developed by Hattox et a!. (1973),
which shows that vanadium atoms in alloys can become magnetic when the
lattice spacing reaches a critical value which is 1.25 times higher than normal
spacing. This approach has successfully explained the magnetic susceptibility
as a function of temperature dependence of bulk samples of AU4 V (Creveling
and Luo, 1968) and Ab V (Creveling and Luo, 1969), as well as of multilayer
films of Au and V (Rau et al., 1989; Nait-Laziz et al., 1993). This
particularity of vanadium may favor the redistribution of charge and spin-
density in its vicinity, causing the appearance of the local moments predicted
by Tugushev (1992) and observed experimentally for Cr-V alloys.
The bulk properties in Cr alloys are extremely sensitive to the presence of
Spin-Density Waves and Charge-Density Waves in Cr Alloys 173

12

f.. . . ! ~aramagnetism
Electronic

3
/ ~':m'''''';'m\\\\\\
.....J .
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7
x (at.%V)

Figure 5. 5 Magnetic phase diagram of the characteristic magnetic field He versus V


concentration. The error bars represent the uncertainty in the He values. and the dashed
line is only a guide to the eye.

impurities. For example, binary Cr alloys doped with Mn (Galkin et al. ,


1996), Fe, and Co, and also ternary Cr(Fe,Co) alloys doped with V (Galkin
et aI., 1997, 1998) or Mn exhibiting thermomagnetic irreversibility with
respect to field-cool ing, below TN and after zero-field cooling (ZFC)
procedure. Fawcett et al. (1999) and Galkin et al. (1998) recently have
shown that for (Cr-2. 7 at. % Fe),-x V x alloy (Fig. 5. 6), there are new
magnetic phases above TN' which are indicated by a magnetic irreversibility
in a characteristic temperature T 1 , as shown in Fig. 5.6. In these alloys when
V concentration is increased, TN decreases and becomes zero for x = 2. 5,
indicating that the long-range order of SOW disappears completely.
Nevertheless, T 1 is observed in higher concentrations (at x = 10). The same
behavior is observed for (Cr-3. 2 at. % Co) I-x V x alloys (Galkin et al. , 1997).
The model that can explain the anomalies observed in magnetic
susceptibility measurements, for the field cooling state in CrFeV and CrCoV
alloys, is based on the formation of the local SOW state, as discussed above.
The regions of short range order of SOW of impurities and different kinds of
defects were shown to lead to a self-redistribution of the charge and spin
densities in their vicinity and, as a consequence, to a local rearrangement of
the SOW. For impurities of Fe and Co that present an intrinsic magnetic
moment, the irreversibility is observed in the paramagnetic phase, associated
with the frustration of the magnetic moment of local SOW.
Thermomagnetic irreversibilities are also observed in CrVMn alloys (de
Oliveira et aI., 1999) above the Neel temperature. In this case T 1 does not
present magnetic field dependence. This behavior is not observed in Cr-Mn
alloys, but it is evident in Cr-V alloys below T loc ' Therefore, the introduction
of nonmagnetic impurities leads to the formation of local SOW and as a
consequence, there is the appearance of local magnetic moments in Cr alloys.
174 A. J. A. de Oliveira and P. C. de Camargo

(Cr-2.7at.%Fe)V

100 200 300 400


Temperature (K)
(a)

300
g
[! 200
~.,
0-
S
E 100
]
U
o 468 10
Temperature (K)
(b)

Figure 5.6 Magnetic phase diagram of (Cr-2. 7 at. % Fe)l-x Vx alloys showing the
irreversibility limiting temperature T, (x) and TN (x) Walkin et al. 1998).

In Section 5.4 the formation of spin-glass phase in Cr-Mn alloys as a


consequence of local SOW moments is discussed.

5.3.3 Correlation with Giant Magnetoresistance and Exchange


Bias in Cr Alloys
The history of the investigation of Cr has two periods: The first corresponds to
the investigation of bulk properties, and is linked to the science for
understanding its intriguing itinerant characteristics. A second distinct period
started with the discovery of giant magnetoresistance in Fe/Cr multilayers by
Baibich et al. (1988) and the emphasis on thin films. The antiferromagnetic
coupling was observed between Fe layers separated by a thin Cr spacer, by
Grunberg et al. (1986), and more recently Bodeker et al. (1998) have
observed the coupling of SOW between Fe multilayers. In this case, the
proximity effects of 20 A Fe cap layers on the SOW, in epitaxial Cr (001 )
films, has been determined by neutron scattering. Unlike in bulk Cr where Q
points along (DOl) direction normal to the free (DOl) surface, in the case of
Spin.Density Waves and Charge-Density Waves in Cr Alloys 175

thin films the direction of Q depends dramatically on the film thickness of


chromium t cr . For t cr < 250 Athe SDW propagates out-of-plane with the spins
in the film plane. For tcr> 1000 A the SDW propagates in the film plane with
the spins out-of-plane perpendicular to the in-plane Fe moments.
Despite the very interesting phenomena observed in thin films, there are
many Cr bulk alloys presenting similar behavior. In particular, Brux et al.
(1995) have observed giant magnetoresistance in magnetically inhomogeneous
bulk granular Cr,oo-, Fe, alloys. The largest magnetoresitence has been
observed for the range 10. 7 ~ x ~ 49.0. The origin of the giant
magnetoresistance in bulk Cr-Fe is attributed to the field-induced alignment of
Fe-rich ferromagnetic clusters in a Cr-rich matrix. In this case, the GMR in Cr-
=
Fe collapses just above x 20 at. % , at which long-range ferromagnetic order
sets in. In Cr-Fe-V and Cr-Fe-Mn alloys Somsen et al. (2000) have observed
that the magnetoresistance shows a strong dependence with V and Mn
concentrations. These authors explain this behavior based on a pinning effect
caused by the antiferromagnetic matrix, that would leave the configuration of
the ferromagnetic components little altered in an external magnetic field
resulting in a smaller magnetoresistance than in the case of the absence of AF
coupling. To confirm such a property it would be necessary to study the
microstructure parallel to the variation of antiferromagnetic exchange of the
matrix. However, the authors do not discuss the possibility of the formation of
local spin-density waves around the impurities such as V and Mn.
Another important characteristic of the Cr alloys is the strong magnetic
irreversibility observed in Cr-1. 3 at. %Si, Cr-3. 2 at. %Co and Cr-2. 7 at. %Fe
(Fawcett et al. , 1999). When these samples are field-cooled in high magnetic
from 400 K and the field removed at low temperature, the magnetization
exhibits a typical "noise," as shown in Fig. 5.7.
A possible explanation of this phenomenon is thermal fluctuation, as
observed in electrical resistivity of pure Cr and CrMn alloys thin films (Michel
et al., 1991, 1993). The noise in the latter case appears only when the
sample enters the SDW phase. These fluctuations will be associated with
fluctuation of wave-vector Q of the domains, which accounts for its decreased
intensity at higher temperatures up to TN in the commensurate SDW phase of
Cr Mn. Therefore, magnetic history of the sample is very important in this
case.
From the technological point of view, the recent advances in the
development of spin valve heads have stimulated interest in the exchange bias
effect in antiferromagnetic/ferromagnetic systems. Exchange bias is
associated with the exchange anisotropy created at interface between an
antiferromagnetic and a ferromagnetic material. There are many systems that
present this anisotropy, in particular, in magnetic multilayers. The classical
case of application of this phenomenon is the system Co/Ni/NiO used in read/
write devices, which shows large magnetoresistance and histeresis loop.
However, there are many systems which exhibit exchange bias without clearly
176 A. J. A. de Oliveira and P. C. de Camargo

3.2
~
] 3.0 \..... .....;;;.~~. cr-1.3at.%Si
"
-e
b
2S
~ 2.8

______---.----1
2.6 '--------'-------''--------'-------'
o 100 200 300 400
Temperature (K)
(a)
5.0
~ Cr-3.2at.%Co
.-, 4.8 ::.J.....
~ .. ~....,.-,;,... ~: ...,
J' ~ ~A->t:, t!-
.toE 4.6 .",..~~~.",~.'

~ 4.4
~

4.2, -

0 100 200 300 400


Temperature (K)
(b)
5.0
.......
... , ...... Cr-2.7 at.%Fe
4.8 .~,"
v..:::: <#,
.::4,f..... .
~
E 4.6 ~'Ay
-e"
I
0
4.4 .:. ;',..:: .. ..
..*:.~.I.

2S '1
~
4.2

4.0
200 250 300 350 400
Temperature (K)
(c)

Figure 5. 7 Temperature dependence of the magnetic susceptibility for three Cr alloys


measured in 100 Oe after zero-field cooling (lower data points for ZFC state) and after
cooling from temperature 400 K in a field. H = 50 kOe (upper data points for high-field
cooling state. Galkin et al. 2000).

depending on antiferromagnetic/ferromagnetic interfaces or which have


multiple random interfaces. These are called materials without well-defined
interfaces or materials with competing magnetic interactions, where, due to
the arrangement of the magnetic ions, different areas (or domains) with
antiferromagnetic or ferromagnetic interactions are created. This category
comprises spin glasses and some ferrimagnets. Typical examples of exchange
Spin-Density Waves and Charge-Density Waves in Cr Alloys 177

bias observed in bulk materials are, CuMn, AgMn, NiMn, and AuFe spin-glass
alloys CNogues and Schuller, 1999). In the particular case of Cr alloys,
recently Galkin et al. C2000, 2001) observed that CrCoV exhibit
ferromagnetic hysterisis and an exchange bias effect in bulk samples. This
behavior appears only for samples that were cooled in high magnetic field, of
the order of 50 kOe. The magnetic hysterisis is attributed to switching
moments of neighboring Co atoms and small Co clusters, resulting from a
delicate competing balance of the coupling energies of Co neighbors and Co
moments with the SOW. Since neighboring Co atoms substitute for Cr that in
the SOW phase belong to sub-lattices of opposite spin, coupling with
sufficiently large amplitude SOW give anti-parallel alignment of the moments in
Co pairs.

5. 4 Spin-Glass Phase in Cr-Alloys

Magnetic behavior of spin-glasses systems has been intensively studied by


different magnetic techniques, as dc magnetic susceptibility, thermal
properties, neutrons diffraction, etc. The prototype of the canonical spin-glass
system is Cu-Mn alloys. The spin-glasses systems are characterized by an
unusual magnetic behavior. After zero-field cool ing, the temperature
dependence of the XC T) measured in small magnetic fields, typically of the
order of H ....., 100 Oe, exhibits a low-temperature maximum, T m. The
magnetization M CH) shows nonlinear field dependence with pronounced
hysteresis and strong relaxation effects, such that, the fluctuations induced by
a random Ruderman-Kittel-Kasuya-Yoshida CRKKY) interaction are the
determinant factor to explain this magnetic behavior. Therefore, impurities
distributions are very important in the formation of spin-glass phase CMydosh,
1993).
Neutron scattering experiments in Cu-Mn alloys allow one to associate the
nature of this behavior with the formation of SOW clusters. The satellite
reflections have a rather wide line width as observed at 1/2 Tj CCable et al. ,
1984; Werner, 1990) and a strong inelastic scattering is also detected at the
satellite peak positions even at far higher temperatures than the freezing
temperature. Thus, the dynamical motion and freezing process of the SOW
clusters would play a essential role in the spin-glass-like behavior of this
system. Since the SOW satellite peak position varies with Mn concentration,
the origin of the SOW formation is considered to be a reflection of the parallel
planes of Fermi surfaces in CuMn alloys CHarders and Well, 1973). Pd-Cr
alloys also show spin-glass behavior with correspondent incommensurate
magnetic satellites, which were associated to SOW clusters by Hirano and
Tsunoda C1999).
178 A. J. A. de Oliveira and P. C. de Camargo

When atoms carrying a magnetic moment are introduced in an


antiferromagnetic system, for example, Fe in Cr, the spin-glass phase can be
considered as an intermediate state between the antiferromagnetic phase for
low concentration of Fe and the ferromagnetic phase that must develop at high
Fe concentrations. Ishikawa et al. (1965) first observed spin-glass like
behavior characterized by magnetic remanescence at low temperatures for
composition range x = 9 to 15 at. % Fe. Strom-Olsen et al. (1979) have
observed a peak in the zero-field cooling susceptibility, which was used to
determine a spin-glass freezing, T G = 9 K at concentration of 14 at. % Fe,
and a Neel temperature of 65 K. On the other hand, in Cr-2 at. % Mo-
14 at. % Fe, where the addition of Mo in Cr reduces the antiferromagnetic
interaction, the Neel temperature is reduced by a factor 2 (TN = 30 K) and the
freezing temperature increases by about the same factor (T G = 17 K). In both
cases, the marked difference between the temperature dependence of the
magnetic susceptibility for the field-cooled state and for zero-field cooling
provided clear evidence for the existence of a spin-glass state.
Alberts and Lourens (1992) also found the spin-glass transition for the
range of 16 at. % Fe~x~20 at. % Fe, but not in lower concentration (x =
12 at. % Fe and 15 at. % Fe). In fact, in this particular case, commensurate
phase (AFo ) coexists with the spin-glass state, for the range of x = 9-
16 at. % Fe. The spin-glass state disappears for x = 19 at. % Fe. The
antiferromagnetic phase boundary was determined using techniques such as
neutron-diffraction, magnetic susceptibility, and electrical resistivity.
On the other hand, Galkin et al. (1995) have observed this anomalous
behavior of Cr-Mn and Cr-Mn-based SOW alloys system at low temperatures.
In the case of these alloys the magnetic behavior differs fundamentally from a
canonical spin-glass, such as Cu-Mn alloys. The temperature T m of the
maximum in X( T) is almost independent of Mn concentration, in contrast with
the proportionality between T m and the concentration of magnetic impurities,
as in the classical Cu-Mn system. The increase of X ( T) with decreasing
temperature at their saturation was observed previously by Maki and Adachi
(1979), at a fixed temperature for composition ranging from x = 2. 2 -
14 . 7 at. % Mn. However, these authors have used a magnetic field of
2.4 kOe causing a saturation such that the zero-field cooling and field cooling
are undistinguishable.
Although, above TN' Cr-Mn alloys obeys a Curie-Weiss law, below TN'
the alloys with low Mn concentration (x < 6. 2 at. % Mn) exhibit Pauli
paramagnetism down to the temperature of the onset of the anomalous
behavior, which indicates that the local moment on the Mn atom is frozen in the
SOW matrix of Cr. Nevertheless, for higher Mn concentration the Curie-Weiss
law is present above and below TN' which is an indication of the existence of a
Mn moment in the SOW phase. However, a strong decrease of the spin-glass
component of magnetic susceptibility for larger Mn concentrations has been
observed, such that 13 at. % of Mn in Cr, the Curie constant increases by
Spin-Density Waves and Charge-Density Waves in Cr Alloys 179

about an order of magnitude and the spin-glass component of magnetic


susceptibility almost disappears (Galkin et aI., 1996). Figure 5.8 shows the
dependence of X de (T) as a function of Mn concentration.

O.I%Mn 10 0.6%Mn
4.5

eo ~ 8
~ 4.0 E
"DO> 0>
i' 6
S 3.5 o
x
2S ~ 4
~3.0 ~
2 '-----L----'_-:':--cc':-----,J
20 40 60 80 100 20 40 60 80 100
Temperature (K) Temperature (K)
(a) (b)
35
30 1.5% Mn
~ 25
E
"DO> 20
I
o
X 15
~10
~
5

o 20 40 60 80 100 o 20 40 60 80 100
Temperature (K) Temperature (K)
(c) (d)

4.6%Mn 1.3% Si 0.6% Mn


Fe

o o 20 40 60 80 100
Temperature (K)
(f)

Figure 5.8 The temperature dependence of the X (T) determined from the magnetization
in constant magnetic field H = 80 De. for binary CrMn alloys and one ternary CrSiMn alloy.
The ZFC and FC data are shown in each case. An extrapolation to zero X (T) of the ZFC
curve. as illustrated by the dashed line for the x = 4.6% Mn sample. is used to define
the pinning temperature T p (Galkin et a!., 1995).

This unusual behavior has been explained using a model in which the spin-
glasses state is formed, not as a result of frustration of the magnetic impurity
moments through their RKKY interaction as in a typical metallic spin-glasses,
but through the frustration of the moments associated with the SOW of the host
Cr. This model proposed by Galkin et al. (1995) is based on the assumption
180 A. J. A. de Oliveira and P. C. de Camargo

of the existence of a frustration on the boundaries of SOW domains.


The frustration, which is the origin of the spin-glass state, is not
associated with the magnetic moment of the impurity, but it is assumed to
occur on the boundaries of the SOW domains. Below a pinning temperature
called T peach Mn magnetic moment is assumed to freeze in a certain
direction, locking up an SOW domain. The domain boundaries would have
frustrated states for the magnetic moments of the SOW, if for one specific Mn
atom the sublattice moments of the domain boundary have to be aligned
parallel and for an other it would require anti-parallel alignment. The direction
of M at higher temperatures, however, may be either parallel or antiparalell
to SOW moment, the latter corresponding to a metastable state.
The canonical metallic spin-glasses Cu-Mn also support an SOW (Cable
et al., 1984; Werner, 1990), but the frustration of the Mn moments in this
case is associated with the existence of 12 wave-vector Q-domains (Cable et
aI., 1982). Therefore, there are some important differences between the
behavior of CrMn alloys and Cu-Mn.
Ternary CrMnV alloys also exhibit spin-glass behavior, as it has been
shown for Cr-1 at. % V-1 at. % Mn by de Oliveira et al. (1999). Those authors
have observed that the introduction of V in these alloys do not change the
pinning temperature, similarly to CrMn.

5.5 Neutrons and X-Ray Scattering: CDW/SW and


SDW Domains

5.5.1 Technical Antiferromagnetism


Technical ferromagnetism is a very well established field of investigation;
however, in the case of antiferromagnetic materials this is hardly mentioned.
The reasons are likely to be the traditional lack of technological interest
of antiferromagnetic materials and the difficulties to characterize
antiferromagnetic domains, which have no net magnetic moment. However,
with the increasing interest in nanotechnology, surfaces and interfaces at
atomic level playa fundamental role and the net magnetic moments at surfaces
and interfaces become a relevant subject. In order to understand the
magnetism of surfaces and interfaces and how is the interlayer coupling, it is
frequently necessary to know what kind of antiferromagnetic domains are
formed and how they can be modified.
In the case of chromium and its antiferromagnetic alloys, as it is
described in previous sections, the magnetic domains are characterized by a
periodicity 0 1 and a sub-lattice magnetization. Therefore, the AF 1 phase is
Spin-Density Waves and Charge-Density Waves in Cr Alloys 181

characterized by domains with two orthogonal directions defined by (OJ' Sj)'


Due to the bcc symmetry of paramagnetic chromium, one shall expect that all
six combinations of OJ and Sj are equally distributed through the volume of the
sample. For samples with good crystalline quality, the Q domains are known
to be of the order of millimeters, The question of the size of S domains is more
difficult to solve because it would require neutrons section topography or non-
resonant magnetic scattering (NRXMS) with azimuthal and translation scans,
as will be discussed in this section. Recently, spin resolved spectroscopy,
using spin-polarized scanning tunneling spectroscopy (SP-STS) has
demonstrated the alternating magnetization of monatomic steps on a Cr (001 )
face. This result was obtained using a Fe coated tip scanning tunneling

microscopy (STM) in a constant current mode, comparing the conductivity :~


for parallel and for (anti) parallel relative spin-orientation between tip and
sample (Kleiber et ai, 2000).
Despite the apparent interest of technical antiferromagnetism for thin films
technology, experimental results (Zabel, 1999) have shown that atomic level
steps and interlayer exchange interactions play the major role, so that domain
boundary energies are easily overcome by these interactions.
However, it is worth mentioning that S domains require much less energy
to be rotated and to be pinned than Q domains, so that, it is likely that without
any external anisotropy, the domains sizes are determined by local defects as
dislocations, vacancies or surface, and interfaces discontinuities.
The question of the existence of an intrinsic COW as a result of a strain
wave caused by the formation of the SOW requires the support of models that
can be divided in two categories. The first is based on the ideas of Young and
Sokoloff (1974) which pointed out that a second-harmonic charge-density-
wave should accompany the incommensurate SOW through electron-phonon
interactions which generates a periodic lattice distortion or strain wave. This
model was further explored by Fishman et al. (1992) who found that a linear
relation between the integrated intensity of the COW satellite and the square of
the integrated intensity of the SOW satellite. The angular coefficient of these
straight lines is related to the Coulomb coupling and increases for Cr-Fe and
Cr-Si alloys, decreasing for Cr-Re, Cr-Mn and Cr-Ru. Based on Fishman
et al. (1992) as an alloy approaches the commensurability the intrinsic COW
increases; therefore, a systematic analysis of the ratio between COW and
SOW intensities for different alloys, would allow one to conclude if there is a
intrinsic COW or if it is only the strain wave caused by the harmonic distortion
of the lattice.
A strain wave is detected either by neutrons or by X-rays; however, an
intrinsic COW would be detected only by X-rays. Nevertheless, it is expected
and all experimental evidences lead to the co-existence of both features,
182 A. J. A. de Oliveira and P. C. de Camargo

which makes interpretation strongly dependent on the model used, and one
must be careful with simplistic approaches that conclude that there is no COW.
The most recent effort to establish the existence of COW was performed
by Strempfer et al. (2000) using high-energy X-rays and measuring the Q-
dependence of the COW second and fourth harmonics, which lead to the
conclusion of the inexistence of intrinsic COW for the model used. The authors
are careful in pointing out that their result is strongly dependent on the model
used for the charge form factor.

5.5.2 Neutrons and X-Ray Two Complementary Techniques Used


in Studies of Cr Magnetism
Neutron techniques are well established for the investigation of magnetic
properties and are especially appropriate for studies of antiferromagnetism.
The excellent review of Fawcett (1988) covers the most important aspects,
but mentions only briefly results on polarized neutrons and on thin films.
Polarized neutrons are a subject that deserves a specific review and will not be
discussed here. The thin film aspect has been quite well covered by Zabel
(1999), who presents a detailed review concerning the investigation of
chromium SOW and COW in films and its coupling with diamagnetic,
paramagnetic and ferromagnetic cap layers. However, X-rays polarization
analysis and some peculiarities of CDW/SW (strain waves) are not treated in
those reviews and are the focus of this section.
Here only a very brief mention of the features concerning neutron
scattering is given, and it is suggested to the interested reader to look at
specific literature of Squires (1978) and the reviews mentioned above. In the
following sections more detailed results are shown and discussed concerning X-
ray magnetic scattering and aspects of CDW/SW and its relevance for the
understanding of magnetic domains.
Thermal neutrons have wavelengths in the range O. 5 - 5. 0 A,
corresponding to energies of 3 to 300 meV, which matches typical interatomic
separations in condensed matter systems and energies of magnetic
excitations. The magnetic moment carried by neutrons interacts with the
spatial variation of the magnetization in materials on the atomic scale. Thus,
they can provide information on magnetic structures and excitations. Distinctly
for X-rays, the cross-sections for the neutron scattering from chemical and
magnetic structures are of similar magnitudes.
By using neutrons one can probe samples inside special environments that
are almost transparent for neutrons, such as furnaces, high-pressure cells and
low-temperatures cryostats. From elastic neutron scattering one can learn
about static and dynamic magnetic ordering, magnitude and orientation of
magnetic moments.
Neutron diffraction form factor carries out information of both orbital and
Spin-Density Waves and Charge-Density Waves in Cr Alloys 183

spin contributions, so that, provided one has the right theoretical model,
orbital moment (L) and spin (S) contributions can be determined. However,
this procedure has led to wrong estimates of the (L) and (S) contribution for
chromium (Stassis et aI., 1975; Strempfer et aI., 1999). A combination of
elastic and inelastic neutron scattering is a unique tool to understand magnetic
order-disorder process and the nature of magnetic fluctuations as described by
Fawcett (1988).
X-ray magnetic scattering is limited by the fact that the magnetic cross
section for photons is reduced by a factor (":;;2 )2as compared to nuclear
charge scattering or to neutrons scattering, which corresponds to a reduction
of 4 x 10- 4 for hv being about 10 keV.
One advantage of high-energy X-ray diffraction is the volume enhancement
gain, due to the small absorption allowing the use of large samples in
transmission mode. In addition to neutrons and X-rays combined techniques,
there is the non-resonant magnetic diffraction at energy of a few keV, that
penetrates only a few micrometers, but allows a unique and not model-
dependent determination of Land S contributions. This is discussed in
Section 5. 5. 3 together with azimuth scans.

5.5.3 Aspects of Neutron and X-Ray Scattering in Cr and C.-Alloys


The preparation of Cr alloys is an interesting procedure to investigate the
changes in magnetic properties due to changes in the electronic structure. The
pioneer work of Koehler et al. (1966), based on neutron and on electronic
transport, has shown that rigid band ideas are well suited for understanding of
the effects of alloying on the Neel temperature, on the sub-lattice magnetic
moment and on the commensurate, incommensurate nature of the magnetic
transition.
This section concentrates on chromium and on chromium-vanadium alloys,
which are incommensurate and present AF2 phase at low temperature. Fawcett
et al. (1994) have presented a broad general review showing many aspects of
the alloying effects.
From the point view of neutrons techniques, there is not much to add to
the recent rev iew by Zabel (1999); however, the new synchrotron sources of
X-rays have brought a few new insights that are reviewed here.
The first, high-resolution X-ray scattering of chromium, where the
magnetism comes only from itinerant electrons was performed by Hill et al.
(1995). The authors have determined the reciprocal space representation of
antiferromagnetic phase (AFM) Cr, showing nuclear peaks, and the COW/SW
and SOW satellites, directly from X-ray measurements. Their representation
of the reciprocal space has no new features, but it was obtained directly from
X-rays. Those authors have not used X-ray polarization analysis. The use of
184 A. J. A. de Oliveira and P. C. de Camargo

high brilliance synchrotron is a fundamental tool for X-ray magnetic scattering


and its polarized nature extends its usefulness as it is shown in Section 5. 5. 4.
There are several formalisms to treat the separation of orbital (L) and
spin (8) contributions; however, an excellent didactic introduction is given by
Brunei and de Bergevin (1991). For more specific treatment the reader is
referred to Gibbs et al. (1990 and Langridge et al. (1997).
An interesting approach for the determination of orbital and spin
contributions was adopted by Strempfer et al. ( 2000 ). The authors have
measured magnetic diffraction of high-energy X-rays photons, combined with
neutrons diffraction in order to find the orbital (L) and spin (8) contributions to
the magnetism of chromium. Taking advantage of the polarization dependence
of the high-energy X-ray cross section, one can determine the polarization of
magnetic moment even without polarization analysis. The differential
scattering cross section for high-energy X-rays can be represented as

da = r2 ( Ad
C ) 2 I S 12 (5.8)
dO 0 1-

where ro is the classical electron radius, Ac is the Compton wavelength, d is


the lattice spacing, and S 1- is the Fourier transform of the spin component
perpendicular to the diffraction plane.
Neutron diffraction intensity is proportional to L (Q) + 28 (Q) of both
orbital L and spin 8 moment and high-energy photons have a cross section that
depends on the Fourier transform of the spin component (8) perpendicular to
the diffraction plane. Therefore, one can also have an indirect estimate of
orbital and spin contribution to the magnetic form factor (Stassis et al.,
1975), using polarized neutrons and choosing the correct model to determine
the form-factor of the induced magnetic moment. More recently, Strempfer
et al. (2000) added new data concluding the fact that the induced form factor
could be described by 40 % spin and 60 % orbital contributions. However, by
comparing the results to an atomic Hartree-Fock calculation, the form factor of
the ordered moment of chromium could be described by pure spin dependence
with no orbital contribution (Strempfer et al. , 2000). This is a clear indication
that the result is totally dependent of the model used and alternative ways of
determining the orbital and spin contribution must be adopted.
Some advantages of high-energy X-rays diffraction are the volume
enhancement gain due to the small absorption allowing the use of large
samples in transmission mode and the access to wide reciprocal space scans,
including higher-order satellites.
From another perspective there is the non-resonant magnetic diffraction at
energy of a few keV that penetrates only a few micrometers, but that allows a
unique determination of Land 8. This aspect will be discussed in the following
section.
Spin-Density Waves and Charge-Density Waves in Cr Alloys 185

5.5.4 Orbital Moment in Cr and Resonant Scattering


Bragg magnetic diffraction or non-resonant X-ray magnetic scattering
(NRXMS), meaning frequencies far from absorption edges of the material,
requires samples of high crystal quality, even when using third generation
synchrotron sources, because of the low magnetic cross section of X-rays.
Chromium and its incommensurate antiferromagnetic alloys are natural
candidates for such measurements because of the isolated magnetic satellites
as described in the introductory section.
The basis for the so-called non-resonant Bragg magnetic scattering was
laid by de Bergevin and Brunei (1970, 1981), and it is based on four
scattering mechanisms for an electromagnetic wave interacting with matter
based on the Thompson model.
The first mechanism would be the electric field of the electromagnetic
radiation interacting with the electron charge causing an electronic dipolar re-
radiation. The second corresponds to the displacement current of the electric
field interacting with the magnetic moment that can have orbital (L) or intrinsic
spin (S) origin, therefore having a quadrupolar magnetic re-radiation. The
third is the gradient of the radiation magnetic field interacting with the material
magnetic moment, accelerating the electronic charge and re-radiating as an
electric dipole. The fourth is the torque caused by the electromagnetic
magnetic field on the material magnetic moment causing a magnetic dipolar re-
radiation.
Except for the first process, all others strongly depend on polarization of
the radiation with respect to the magnetic moment in the material.
S. S. 4.1 Non-Resonant X-Ray Magnetic Scattering
Now a specific experiment with a high quality Cr crystal is described as
performed by Mannix et al. (2001). The components of the orbital moment
(L) and spin (S) magnetic moments in the NRXMS regime are defined within
the standard coordinate system in Fig. 5. 9 as L I and S I ( i = 1,2, 3). For
incident a polarized photons, the NRXMS scattering amplitude is (de Bergevin
and Brunei, 1981):

f u-u ]
f non-res = [ f u-"
sin2 eS 2 (Q) }
{ 2sin2 ecose[L, (Q) + S 1 (Q) ] + [sin2 eS 3 (Q) ]
(5.9)

where hw is the incident energy and '0


is the classical electron radius. The
different polarization dependence of Land S components, leads to the
possibility of spin and orbit moment separation using polarization analysis
186 A. J. A. de Oliveira and P. C. de Camargo

(Gibbs et a!., 1991; Langridge et a!., 1997).

Polarisation analysis
Azimuthal <P

C1~
Q (001+0):
I k'
I k
G"
...............................................'-'>.""'--,

Figure 5. 9 Incident a-polarized photons with wave vector k, are diffracted from the
sample with scattered wave vector k'. The specular (001 + 6) spin density wave was
studied in both non-resonant and resonant scattering regimes. The sample is aligned with
the crystal four-fold axis along parallel to the scattering plane which is defined as azimuthal
C1> = O. At this value of C1>. the S x polarization domain has magnetic moments parallel to
U, and the Sy domain has magnetic moments directed along U 2 The standard axes U,
( ; = 1 .2.3) are discussed in the text. The magnetic scattering from the two polarization
domains is averaged by azimuthal C1> (Mannix et al. 2001).

The magnetization components L{ and S/' in Eq. (5.9), can be expressed


in terms of the azimuthal-a> rotation angle as

S, (Q) = S(Q)sina> }
L,(Q) = L(Q)sina> (5.10)
S2 (Q) = S (Q)cosa>
Here, we have assumed that Land S are co-linear and the formation of a
linear SOW. The LiS ratio, for the azimuthal averaged intensities, follows
from Eq. (5.9) as

(5. 11)

1~C7 and a>: are the a> averaged intensities, corrected by the different analyzer
rocking curve widths for the aa and a1r polarization. An azimuthal dependence
is expected for a linear SOW because the moments are oriented ninety
degrees from each other in the basal plane for this structure. Therefore, in the
AF, phase, the azimuthal intensities give unique insight into the nature of the
chromium spin-density wave, directly confirming if it is linear from an observed
periodicity. In the AF 2 phase, the polarization flips parallel to Q and no
magnetization components are rotated in ep around the scattering vector.
Spin-Density Waves and Charge-Density Waves in Cr Alloys 187

Consequently, no azimuthal dependence is expected in this phase.


Only (001 () SOW magnetic reflections are accessible at 5.219 keY,
due to the relatively small lattice parameter (ao = 2.88 A, implying in a rather
large reciprocal space. Radial scans normal to the surface (parallel to Q) of
the (001 () spin density wave reflection, are compared in Fig. 5.10. The
magnetic peaks are resolution limited, so that it is not possible to extract an
exact correlation length for the S domain from the widths. However, a lower
limit of -4000 A can be deduced. A larger background for the eUT polarization
arises from the specular ridge and this is accounted for in the Lorentzian fits to
the data. The aa and alT scattered polarizations were collected for several
azimuthal angles (<l and their integrated intensities are shown in Fig. 5. 9.
The open circles are for the aa intensities and the solid circles for the alT, in
the AF 1 phase (T = 140 K). This figure demonstrates that chromium indeed
forms a linear spin-density wave because of the observed azimuthal
periodicity. The intensities exhibit a two-fold symmetry, consistent with an
almost entirely single polarization domain population. Since the intensities do
not go to zero at <l> = 90, the scattering volume cannot be completely single S
domain (Eq. (5.10. Domain volume fractions, v y = 0.83 0.03 and v x =
O. 17 0 . 03, of the 8 y and 8 x polarization domains, have been deduced from
Fig. 5.9. The size of the S domains could have been further investigated by
translating the sample in the X-ray beam. However, this kind of analysis was
not undertaken. The solid line is a fit using Eq. (5.10) and gives the aa and
alT intensities averaged across the v x and v y domain volumes. These values
are marked with an arrow on the left hand axis and correspond to 1'{p0" = (5.7
0.1) x 10- 6 and 1~1T=(1.080.06) x 10- 6 . From Eq. (5.11), this gives a
ratio of L / 8 = O. 008 O. 03. The orbital magnetization in chromium is
therefore zero (less than 1% ) . To the accuracy of these measurements, an
upper limit of 3 % orbital density can be deduced. This negligible value of <U
is expected for 3d 5 electronic ground state of chromium. The solid triangles in
Fig. 5. 11 are the intensities alT in the AF2 phase (T = 50 K), the weaker
intensity arises from the geometrical terms for 8 3 in Eq. (5.9). No azimuthal
dependence is observed in this phase as anticipated for S II Q. The intensities
aa are zero below T SF' because the magnetization components along 8 3 now
parallel to the scattering vector. This also implies that a single Q domain is
being probed, because otherwise one could have azimuthal dependence due to
8 1 or 8 2 contributions.
From this direct measurement of zero orbital magnetization in chromium,
we discuss in Section 5.5.4.2 the resonant X-ray magnetic scattering (RXMS)
by considering both the non-resonant and resonant amplitudes. For this
purpose, it is convenient to write the spin only (L = 0) non-resonant amplitude
from Eq. (5.9) as

tnoores = [ t::ores ] . -hw ro 2Sf


t:;;.,es =- I me 2 m
(Q) smB
. Z2 cosB
x [ smB(zlcosB
. .
+ z3 sm B)
]
'
(5.12)
188 A. J. A. de Oliveira and P. C. de Camargo

10- 1

?0
10-2
.:
0: rr-rr
rr-1t

~
0
.iii 10-3
~"
..=
10-4

1.035 1.045 1.055


L (rlu)

Figure 5. 10 Typical L-scans in the non-resonant regime used to integrate the scattered
intensities in the (J(J and (Jll' channels. The data are on a logarithmic scale and the larger
background for the (J(J polarization is from the specular charge rod. The solid lines are
Lorentzian fits to the data (Mannix et al. , 2001).

1.0 00 o

of' 0.8
o
rrrr TSDW (T=140 K)
rr1t TSDW (7'=140 K)
o rr1t LSDW (T=50 K)
.~ 0.4
.s'"
.
o 0.51t 1t 1.51t
Azimuthcl q, (Radians)

Figure 5.11 The azimuthal dependence of the (J(J (open circles) and (Jll' (solid circles)
non-resonant scattering in the transverse spin density wave phase (T= 140 K). The solid
line is a fit, giving LiS = O. The data give unique evidence that chromium forms a linear
polarized spin density wave. The solid triangles are data taken at T=50 K, in longitudinal
spin density wave phase (S II Q). No azimuthal dependence is expected in this phase
(Mannix et al. , 2001).

Here, 25 is the ordered magnetic moment, fmCQ) is the spin only form factor
and ZI.2.3 are unit vectors along the magnetization direction V 12 3 shown in
Fig. 5.9. A pre-factor of I f noores I =2SChwlmc 2 ) ro~5 x 10- 3 ro can be
determined to the geometrical terms in Eq. C5. 12). This value is then used in
the models of the resonant scattering described in the next section.
Spin-Density Waves and Charge-Density Waves in Cr Alloys 189

S. S. 4. 2 Resonant X-Ray Magnetic Scattering


The resonant X-ray magnetic scattering ( RXMS) is usually associated to
dipolar or quadrupolar transitions associated to one of the absorption edges of
the element of interest. In the case of K absorption edge of chromium, Hill
et al. (1995) have found no resonance; however, it will be shown in this
section that there is a clear enhancement of the magnetic scattering intensity,
consistent with the assumption that it comes from transitions of core 1s level
electrons to 4p ~tates. The quadrupolar transitions to the magnetic 4d states
are essentially non-existent due to the absence of orbital contribution as it was
shown in the previous section.
The RXMS dipole and quadrupole amplitudes are conveniently given by
Hill and McMorrow (1996). For incident a-polarized photons the dipole
resonant magnetic scattering amplitude is

resl -
f El -
[mJ-
f El - [(E. - Eo
A E1
-hWE1)/(1/2) _ i] x
[ .
Z3 sm
e- ZlCOS
e]
(5. 13)
where A EI determines the strength of the magnetic scattering (at the magnetic
wave-vector) and is related to atomic overlap integrals between initial and
excited states. The denominator describes the deviation from resonance in
units of the core hole lifetime'" and accounts for the Lorentzian profile of the
resonance. E. is the energy of the initial state and Eo is the energy of the
excited intermediate state. Typical values of A EI for transition metals are
about 10- 2 ro (Stunault et al. , 1999). The total magnetic scattering amplitude
is the sum of the non-resonant (Eq. (5. 12 and resonant (Eq. (5. 13
contributions. The absorption coefficients in the vicinity of the chromium K
absorption edge, extracted from fluorescence spectra (Eisebitt et al. , 1993),
are shown in Fig. 5. 12. The sol id Iine is a fit to the data, consisting of an
arctangent function to describe the absorption step function and a Lorentzian
function in the white line region. Additional peaks are later added to account
for the EXAFS oscillations after the edge. By careful fitting of the pre-edge
tail, a value of the core-hole lifetime of ,.. =
(1. 1 0.5) eV was obtained (de
Bergevin et aI., 1998). The measured RXMS peaks are slightly broader than
this due to finite energy resolution effects. The integrated intensities of
the (001 + 5) spin density wave, in the vicinity of the chromium Kedge
(E = 5.989 keV) , taken at an azimuthal angle ([> = 0, are shown in
Fig. 5.12b-d. The open circles in Fig. 5. 12b are for the aa scattered
polarization in the AF 1 phase (T =
140 K) phase and the corresponding an
polarization are the solid circles in Fig. 5.12d. Solid triangles in Fig. 5. 12d
=
represent the an intensities in the AF 2 phase (T 50 K). These data have not
been corrected for absorption, but corrections have been made for cross-talk
(Vaillant, 1977), due to the deviation from 45 of the analyzer crystal Bragg
angle. RXMS is observed. The peak appears only for the an scattered
190 A. J. A. de Oliveira and P. C. de Camargo

polarization and occurs in the white line energy region. These are strong
indications that the resonance arises from dipole transitions to the 4p states.
Quadrupole resonant magnetic scattering depends on orbital moment
interactions and is not expected in this experiment. The experimental result is
quite well described considering only dipolar contribution agrees with the
results shown in the previous section and with the theory for zero orbital
moment contribution for the magnetism of chromium.
To describe the energy profiles in Fig. 5.12b-d, we have initially
considered only one dipole resonant magnetic amplitude and the interference
from the non-resonant magnetic amplitude. The total magnetic scattering
intensities for the aa and an polarization in the different temperature phases
of chromium are

(5. 14)
i=x,y

'111" (m) -
1AF 1 '" -
A "" V
SF/~Y I
I [f"'"
nonres
(m .)
",,I
+ [E A E1 [- ZI (CD, Dcose]
- E c -hWE1/(,/2) - i]
] 1
2

(5.15)
I AF2 =0 (5.16)

I"'" -
AF 1 - A SF
I[ f ",.
nonres
[E. - E c
A E1 [Z3 sin e]
- hWEl / ( , /2) - i]
] I (5.17)

Since the measured intensities are not on an absolute scale, the scale
factor A SF is used. The V I ( i = x , y) refers to the volume fractions of the x and
y domains, present in the AF 1 phase. The scattered intensity from these
domain volumes depends on the azimuthal angle CD (see Eqs. (5. 14) and
(5. 15)). At CD = 0, the scattering from the two domains may be separated
with polarization analysis, provided that L = O. In the AF 2 phase, there can
only be one domain volume, since S II Q. The scattered intensity does not
depend on the azimuthal angle below T SF ( Eq. (5. 17)) and the incident s-
polarized photons are scattered with a change of polarization (Eq. (5. 16)).
The solid lines in Fig. 5. 12b - d are fits using Eqs. (5. 14) - (5.17). The
chromium crystal had an extended face and the change in absorption across the
K edge region was accounted for by dividing the calculated integrated
intensities by the absorption coefficients (Fig. 5. 12a). In our models, we
have assumed that the domain fractions do not change as a function of photon
penetration depth. The relative domain volumes deduced from the azimuthal
dependence in Section 5. 3 V Y = O. 87 and V x = O. 13 were used in the fits to
data in the AF, phase (Fig. 5.12b and 5.12c). In order to model these data,
f noores was kept constant at its calculated value of 5.0 x 10 -3 ro, with the
geometrical factors in Eq. (5.12). For the aa intensities, in Fig. 5. 12b, the
scale factor is the only variable. The solid line fit is in very good agreement
with the data and shows that the energy profile arises from the change in
absorption for this polarization. Note, that this figure is equivalent to Fig. 5.5
presented by Hill et al. (1995), measured without polarization analysis and for
Spin-Density Waves and Charge-Density Waves in Cr Alloys 191

4000
3200
0: 11 (em-I)
E2400
~

-;: 1600
E2- -EI
800
0
5.92 5.94 5.96 5.98 6.00 6.02 6.04
Energy (keY)
(a)
.
~

0
I
2

X
~

P
Vi
<::
(l)

.e
.5
0
5.92 5.94 5.96 5.98 6.00 6.02 6.04
Energy (keY)
(b)

~ 2.50
"i'
0

x
~
.: a1t TSDW (T=140 K)
.~ 1.25
<::
B
.=:
o
5.92 5.94 5.96 5.98 6.00 6.02 6.04
Energy (keY)
(e)

f" 2.50
o
2S ~: a1t LSDW (7'=50 K)
p
.~ 1.25

.~
o
5.92 5.94 5.96 5.98 6.00 6.02 6.04
Energy (keY)
(d)

Figure 5. 12 (a) Absorption coefficients obtained from the fluorescence taken in the vicinity
of the K absorption edge; (b) The aa resonant magnetic scattering in the AF, phase;
(c)The all" resonant scattering intensity in the AF, phase; (d)The all" resonant scattering in
AF2 phase. The solid lines are fits using the models described in the text (Mannix et al. ,
2000.
192 A. J. A. de Oliveira and P. C. de Camargo

which resonant magnetic scattering is not observed. The obtained value of


A SF = 1. 1 0.2, was subsequently used to model the AF 1 em data in Fig. 5. 12c. In
this fit to the data, the variables were A E1 , hWEl and r, which were found to
be A E1 = (3.2 O. 2) x 10- 3 ro, hWEl = (5.989 0.001) keY and r = (2. 4
0.5) eV. The observed r is significantly broader than that deduced from the
fluorescence (1. 1 eV). This broadening reflects the band character of the p
states in chromium. The values of A E1 , hWEl and r obtained, were then used
in the fit to the an AF 1 data, in Fig. 5. 12d. The domain volume fraction was
assumed to be one and only variable was the scale factor, which was found to
be A SF = O. 9 O. 2. This value is in very good agreement with the one
deduced in the AF 1 phase, in good accord with the expectation of a single
polarization state in this phase. This value also suggests that the (001) face of
chromium remains single-Q across this temperature range. The apparent shift
in position and shape if the an intensities of Figs. 5. 12c and 5. 12d, arises
from a change of sign in the interference between non-resonant and resonant
scattering amplitudes. This is described by the Zl and Z3 components in the
dipole cross-section (Eq. (5.13. The fits to the energy profiles are slightly
improved by allowing for a small quadrupole amplitude of 3 x 10- 4 ro and
energy hWE2 = 5. 985 keV. This gives an upper limit to the quadrupole
amplitude of at least one order of magnitude smaller than that of the dipole.
Similar weak quadrupole amplitudes have been determined for other 3d5
compounds, RbMnF 3 (Stunault et ai., 1999) and MnO (Neubeck, 2000). The
values for the dipole and quadrupole amplitudes, found from our models of the
RXMS in chromium, are consistent with those derived by Renninger (1937)
and Hill and McMorrow (1996), taking into account the smaller magnetic
moment of chromium. The agreement together with the good fit to the data in
Fig. 5. 12, gives confidence in the validity of this approach and to the
quadrupole model developed by Neubeck et al. (2001).
These investigations using X-ray magnetic scattering with polarization
analysis, give a fresh and unequivocal insight into the fascinating magnetic
properties of chromium. The small X-ray beam size individually probes the
relatively large Q domains and our NRXMS investigations give direct evidence
for the formation of two polarization S domains in the AF 1 phase. The models
used in the resonant scattering regime also support the formation of two
polarization domains in the AF 1 phase and a single domain in the AF 2 phase.
The NRXMS azimuthal dependence provides unique evidence that chromium
forms a linear polarized spin density wave. We have subsequently used these
intensities to directly determine the zero orbital magnetic moment in
chromium. This value of L = 0 is expected for the chromium electronic
configuration. We have reported on resonant X-ray magnetic scattering from
chromium, using polarization analysis, for the first time. A consistent model of
the energy profiles has been derived in this energy regime, from the resonant
and non-resonant scattering amplitudes. The scattering arises primarily from
dipole transitions to the 4p states. Since the 3d states are of most interest to
Spin-Density Waves and Charge-Density Waves in Cr Alloys 193

the understanding the magnetism of chromium, it is natural that the band


structure calculations have focused mainly on the 3d band and little work has
been published on the 4p spin polarization. However, further calculations of
the 4p density of states are needed to provide a more quantitative model of the
dipole RXMS described in this paper. RXMS involving quadrupole transitions
to the 3d band, has not been observed. This effect can be understood from
recent models of the role of the orbital moment in 3d K-edge quadrupole
resonance. For the specific case of the 3d5 ion, the scattering amplitude is
zero. The derived resonant and non-resonant amplitudes from our model are
completely consistent with values for other 3d5 systems, for example RbMnF3
and MnO. Dipole transitions, probing the chromium 3d states, are allowed at
the spin orbit split L absorption edges of chromium.

5.5.5 Magnetic Q Domains in Cr (001) Face


Domains within an antiferromagnetic sample are defined naturally either by
internal defects, as a result of the kinetics of the phase transition or else by
external anisotropy, due to applied pressure or magnetic field. In the case of
chromium, the magnetic domains have been determined with neutron
diffraction. There has been ongoing controversy about the origin and nature of
the magnetic domain states since the 1960s (Werner et aI., 1966) and the
subject still remains an open question (Hill et aI., 1995). On the other hand,
the comprehension of the nature of the domains occurring near the surface of
chromium is of great importance, especially because it is related to how spins
are distributed in chromium films with different crystalline free face
orientations. In case of antiferromagnetic alloys of chromium, because the
magnetic modulation propagates along one of the fourfold axes of the cubic
structure, there is the possibility of three different magnetic Q domains. Ando
and Hosoya, (1978) in a pioneering neutron topography experiment, have
roughly determined the domains in high-quality chromium crystal. In order to
image the domains they have used the magnetic satellite Bragg peaks, which
occur around each nuclear (DOl) prohibited reflection. Each of these satellites
is associated with one type of domain. However, the magnetic structure is
associated also with the charge-density waves/strain waves (CDW/SW) ,
producing a distortion, also incommensurate with the crystal lattice, with a
period twice that of the magnetic structure. The corresponding CDW/SW
satellites are consequently located around the usual Bragg peak, but at a
distance that can be easily resolved.
A practical rule to associate the domains with reciprocal space can be
written as follows: the domain pair, (0 I , S j) i =1= j, is readily associated with
the reciprocal space. Choosing one of the axes H, K or L, and the direction
of one 0+/- satellite pair that defines Q direction, then the spin S must be
oriented in that direction that does not involve the axes already considered or
194 A. J. A. de Oliveira and P. C. de Camargo

the direction along the Q+I- satellites pair. Obviously, if the pair Q+I- is
along one of the reciprocal space axes, then there are two possibilities for the
spin in the AF, phase. For the AF 2 phase the spin is obviously defined by the
Q+I- satellite direction.
The intensity of these SW/COW satellites was measured to be 8 x 10 4
times smaller than the corresponding Bragg peak. Recently, de Camargo
et al. (2001a,2001b) and Baruchel et al. (2001) have shown that the X-ray
topography of a given COW/SW satellite reflection provides a direct image of
the Q domains with orientation corresponding to that satellite. Combining the
observations of the three independent satellites allows one to obtain the entire
Q domain structure of the crystal.
Ando and Hosoya (1978) have pointed out that for pure chromium, "even
without any external imposed anisotropy, some crystals may present marked
preference to have Q along one particular cubic axis". Following X-ray
topography de Camargo et al. (2001 a, 2000b) have shown that at (001) free
faces of chromium, a single-Q state is formed, as shown in Fig. 5. 13, which
is an X-ray reflection topography of the (0,0,2+26) satellite performed at 10-
19 at the European Synchrotron Radiation Facility (ESRF). The whole larger
surface of the L-shaped face was illuminated with energy of 25 keY and beam
divergence of 10 arcsec. Under these experimental conditions only Q domains
along (001) direction will appear, so that the L-shaped surface COW
diffraction picture is an evidence of the single Q nature of that face. The
contrasting region near the L external side corresponds to a defect that is also
detected on the regular Bragg topography (Baruchel et al. , 2001), and it is
associated to growth band defects.

Figure 5. 13 A white beam 200 topograph, in transmission geometry, of this sample: the
dark lines marked with arrows correspond to growth bands, observed also in section
topography experiment and not shown here. Figure 5. 12b refers to X-ray topography of
the COW satellite (0, 0,2+25), of sample 3, showing that the surface is mostly single
Q. The energy used was 25 keV with a divergence of 10 arcsec illuminating the whole
sample surface with the (001) surface inclined 9.3 and the FReLoN camera perpendicular
to the incident beam (de Camargo et al. , 2001a).

The evidence of the single-Q state in (001) face of the chromium was
reported previously by Hill et al. (1995) and de Camargo et al. (1999).
Considering that the size of the domains can reach up to 10 mm3 , one should
not be surprised to find single-Q domains when probing regions near the (001)
surface. In fact, it is expected that a preferred expansion direction shall occur
Spin-Density Waves and Charge-Density Waves in Cr Alloys 195

along Q, pointing perpendicularly to the free surface. An atomically flat free


(001) face is also supposed to be ferromagnetic ordered, but would have in-
plane spins, possibly due to the large demagnetization factor. This speculation
is supported by B6deker et al. (1999), and their experiments with magnetic
and non-magnetic cap layers on chromium (001) face. Those authors have
shown that a 20-A iron cap layer on a 3000-A chromium film is enough to cause
a Q switching, becoming parallel to the (001) face.
The results shown here give new evidence that the single-Q state
observed for (001) surfaces is due to the large defect, which characterizes
free surfaces. In Section 5. 5. 3 the NRXMS azimuthal scan also supports the
idea that Cr (001) face is single Q. As a consequence, one can expect that
the external geometry of a chromium crystal may determine the domain
distribution in its volume and affect other properties like magnetic
susceptibility and magnetoresistance below the Neel transition. This may also
be the case for other antiferromagnetic materials with large magnetic domains.

5.5.6 Paramagnetic Critical Fluctuations in Cr


and in Cr-O. 2 at. % V
Besides the well characterized SDW formed below the Neel temperature, it is
interesting to also observe the well defined critical magnetic scattering in the
paramagnetic regime. This section discusses the paramagnetic critical
fluctuations for both Cr that has a first order Neel transition, and that of
Cr-0.2 at. % V, which has a second order antiferromagnetic-paramagnetic
transition firstly observed by Noakes et al. (1990), just above the Neel
temperature and for energy transfer less than 4 meV. Latter this measurements
were extended to temperatures up to 1. 3 TN and energies in the range of - 4
to - 20 meV confirming previous results.
The combination of elastic and non-elastic scattering can be used to
investigate the nature of critical scattering. The weak first order transition in
chromium shows a discontinuous neutron scattering intensity at TN = 311 K
and is transformed to a second order transition al with addition of as little as
0.2 at. % V, showing a continuous scattering intensity around TN = 289 K
(Noakes et al. , 1990). Therefore, the critical behavior of this alloy can be
compared with that of chromium. The analysis of the dynamical susceptibility
of Cr was contrasted with that of Cr-O . 2 at. % V by Noakes et al. (1990).
Sato and Maki (1974) used the approach to describe the dynamical
susceptibility as a result of two band incommensurate itinerant-electron
antiferromagnet, and the imaginary part of the dynamic susceptibility can be
written as a single expression:
196 A. J. A. de Oliveira and P. C. de Camargo

(xO/r 2)N
X(Q,w) = --....:..:...----- (5. 18)
A 4 [/(2 + R(q)]2 + w 2
where XO is the coefficient of the Curie law susceptibility in the non-interacting
limit, r is the length scale of the magnetic interaction, A is the magnetic
stiffness, and /( is the inverse correlation length, which is temperature
dependent with a critical exponent of 1/2, consistently with a mean field
theory (Noakes et aI., 1990a) and R(q) is the real part of self energy and is
zero at ordering wave vector. The self-energy function R (q) may be wr itten
as

where 5 is the incommensurability parameter.


Performing elastic and inelastic scans of six satellite positions near (00l)
the model described above does not fit the experimental data well, because
the peaks of the observed scattering in constant-energy scans appear to be
moving slightly toward the commensurate position as energy transfer
increases, whereas this model line shape does not. Attempts to correct diffuse
scattering did not improve the fit.
Following Sato and Maki' s idea that the simplest function that describes
R depends on q2, and taking in to consideration the six fold symmetry around
(00l) position, the polynomial expression that involves only squares of q is a
polynomial of order four in q, which involves the squares of the q x' q y and q z
components, where q is the offset from the commensurate wave vector.
Considering Q = (001), then q = 0 and R = 5 2 /4, resulting in an energy of
N . 5 2 14 above that minimum. For the sake of realizing a physical meaning
one shall consider that N . 5 2 14 corresponds to 1. 3 THz, while TN
corresponds to 6. 3 THz. In Cr-O. 2 at. % V, 5 increases and the referred
energy corresponds to 2. 1 THz. The expression for X ( Q , w) therefore has a
intrinsic response to commensurate position, but its magnitude relative to that
at the incipient satellite positions depends on 5 which is a feature intrinsic of
incommensurate systems. As 5 increases, for example, adding vanadium the
susceptibility at the commensurate position shall decrease relative to the peak
at incipient satellite position.
Recently, Stockert et al. (2000) have extended the critical fluctuation
experiment to higher energies and temperatures (400 K) and found that the
Sato-Maki approach describes quite well the magnetic response in
paramagnetic chromium over a wide energy and temperature range, with high
q resolution to resolve the incommensurate fluctuations.
The experimental data up to 400 K can be very well fitted considering that
the response function has an energy scale of the spin fluctuations that varies
linearly with temperature. The spin fluctuation frequency rate of change is
.t..WSF 1.t..K :::::::54 \.leV IK in the temperature interval 313 K < T < 400 K.
Spin-Density Waves and Charge-Density Waves in Cr Alloys 197

In conclusion, it has been shown that critical fluctuations in the


paramagnetic phase are not affected by the order of the phase transition. The
paramagnetic susceptibilities of both Cr and of Cr-O. 2 at. % V alloy, are well
described in terms of the two-band itinerant electron model of the magnetism of
chromium for temperatures well above TN' and for a large energy transfer
range, provided one uses the true six fold around the nuclear reciprocal-lattice
point of the paramagnetic state.

5.6 Ultrasonic Measurements in Cr and Cr-Alloys

The magnetoelasticity of chromium and of its antiferromagnetic alloys are quite


completely reviewed in Fawcett et al. (1994); therefore, only a very brief
comment is included here in order to facilitate the reader's understanding of
the arguments used in this section.
As most antiferromagnetic materials, chromium and its antiferromagnetic
alloys exhibit a negative Gruneisen parameter,

(5.20)

The difficulty to determine the actual values of the magnetovolume near


TN' dw = dV/ V, is the strong sample quality dependence. The connection
between Gruneisen parameter and the bulk volume may be given as

(5.21)

where

B = (V ~ ~)
N TN'
(5.22)

Corrections can be made using the thermal expansion coefficient, but it


seams that a reliable and useful procedure is to define the corresponding
magnetic Gruneisen parameter (Fawcett, 1988; Fawcett et aI., 1994)
1 llB (t)
r, =- B N TN llf3(t) (5.23)

where f3 (t) is the thermal expansion coefficient, and t is the reduced


temperature( T/ TN) .
This definition is being used for chromium and for several alloys indicating
a surprisingly linear dependence of the parameters llB ( t) as a function of
llf3 (t) for both regimes in the ordered phase as well as in the disordered
phase, for many different alloys.
The question of what is the actual thermal expansion coefficient at TN has
198 A. J. A. de Oliveira and P. C. de Camargo

no unique answer, because it depends on crystal quality. The transition is a


first order with a clear a discontinuous change in volume is expected, so that
one can define only the thermal expansion coefficients above and below TN.
For Cr-V alloys II varies from -50 to -30. The parameter II can reach
-600 in the case of Cr-Ge alloys, indicating a strongly nonlinear behavior.
The transition from paramagnetic to AF 1 phase produces a dramatic
change in the longitudinal compression wave velocity corresponding to the
elastic constant C 11 The pronounced sharp minimum observed at TN is in fact
expected, because of the lattice expansion along (001) direction as chromium
becomes antiferromagnetic.
The softening of the C 11 elastic constant starts well above the Neel
temperature and in a temperature interval of less than one degree it shows a
sharp minimum, slowly recovering and overcoming the paramagnetic value.
Measurements in a wide temperature range, represented in a compressed
scale, would clearly show the normal elastic constant trend increasing as the
temperature decreases, with a magnetic anomaly that is associated with the
magneto-volume change near the Neel transition. Comparing chromium with its
antiferromagnetic chromium-vanadium alloys one could subtract what can be
called normal behavior, taking, for example, the paramagnetic Cr-5 at. % V.
Despite the fact that chromium is BCC and only three elastic constants are
enough to define its elastic properties, in doing elastic measurements in the
antiferromagnetic phase it is necessary to consider the orthorhombic symmetry
in the AF 1 phase and tetragonal symmetry in the AF2 phase. These
measurements were performed by Muir et al. (1987), doing ultrasonic
measurements after field cooling a chromium single crystal and obtaining a
state that could be characterized as single Q and single S. These
measurements have shown that the phenomenological theory by Walker
( 1980) would correctly account for the magnitude of magnetovolume
contribution to the elastic properties.
Ultrasonic and thermal expansion measurements were performed on
Cr-02 at. % V single crystal by de Camargo et al. (1988). This is the same
crystal where neutron critical scattering was investigated showing a second
order transition.
Most ultrasonic data on chromium and on its antiferromagnetic alloys have
been explored using single frequency transducers that are solidly attached to
the samples. The bonding characteristics are often difficult to reproduce and
may also cause undesirable stresses. In the case of ultrasonic attenuation
these stresses may represent a real challenge in obtaining reproducible results
for different transducers. Recently, it was demonstrated that it is worthwhile
to use noncontact transducers coupled with a liquid medium (de Camargo
et al. , 2001 b) and this is described in more detail below.
In general, changes in the ultrasonic pulse amplitude and traveling time
are measured, and then the ultrasonic attenuation and elastic constants are
determined for fixed frequencies that are given by the odd harmonics of the
Spin-Density Waves and Charge-Density Waves in Cr Alloys 199

transducer coupled to the sample. It is difficult to reproduce the same


conditions of coupling between transducer and sample, therefore most
analyses are limited to a few fixed frequencies. An investigation of the
ultrasonic pulse propagation close to a magnetic phase transformation, for a
continuous frequency range, is desirable and is presented using wide band
transducers and digital spectral analysis.
The material chosen is a single crystal of Cr-O. 18 at. %, which has a
magnetic phase transformation from the paramagnetic to an antiferromagnetic
state at 318 K (Fawcett et al., 1994; Bosshoff et al., 1993). Similarly to
pure chromium, the ultrasonic attenuation and internal friction for this alloy
have a contribution due to the coupling between the SOW in the
antiferromagnetic phase and the oscillatory elastic strain introduced by the
ultrasonic pulse. In this section, we have shown that the ultrasonic spectral
analysis approach is highly advantageous in many aspects for the investigation
of phase transitions and shall be more intensively used in the case of
magnetoelastic interactions.
Figure 5. 14 shows the typical pulse form in time (Fig. 5. 14a) and
frequency domains (Fig. 5. 14b) as obtained from the wide band transducer.
The continuous line corresponds to the entrance echo and the dotted line refers
to the bottom echo. In Fig. 5. 14b, it is evident that the material under
investigation interacts more strongly with the pulse for frequencies in the
interval from 20 to 40 MHz. Above 40 MHz, the response of the system is
limited by the reception band and both echoes follow roughly the same
pattern. Although some beam diffraction effects are expected, it should occur
preferentially for lower frequencies and seems to be negl igible (Fig. 5. 14b) .
Accurate correction of the diffraction effects for our case is a complex task,
since the transducer is focalized, so multiple echo analysis is to be avoided.

400 475 550 625 700 o 10 20 30 40 50 60 70


Time (ns) Frequency (MHz)
(a) (b)

Figure 5. 14 Echoes in the time (a) and frequency domain (b). The solid lines
correspond to the entrance echo and the dotted lines correspond to the backwall echo
(de Camargo et ai., 2001b).

The spectral analysis of the bottom echo for different temperature regions
corresponding to the paramagnetic phase, transition and antiferromagnetic
200 A. J. A. de Oliveira and P. C. de Camargo

phases is shown in Fig. 5. 15. Three different temperatures for each region are
shown in order to make evident the reproducibility of the results. The
frequency spectrum relative to the paramagnetic (PM) phase has features
similar to the entrance echo (Fig. 5. 14b), however it presents smaller
amplitude around the first peak at the frequencies above 20 MHz. For the
antiferromagnetic phase the general aspect of the frequency spectrum is kept,
but with smaller amplitudes for the whole frequency interval. These results
lead to the conclusion that there is no strong frequency dependence for the
attenuation of the ultrasonic pulses in both paramagnetic and antiferromagnetic
phases; therefore, domain wall movement is not relevant for absorption of
energy in this frequency range. The transition region is just where the
frequency spectrum shows the strongest frequency selected interaction,
causing an almost total absorption at frequencies above 20 MHz, indicating
that the magnetic ordering processes strongly interact with the elastic waves,
possibly magneto-elastically. In Fig. 5. 16 the frequency spectrum as a function
of temperature is presented. The amplitude of the received pulse is shown as a
function of frequency and temperature and different cuts at fixed temperature or
fixed frequencies can be observed. A clear depression of amplitude is
observed around 318 K for the higher frequencies. It can be seen that there is
a narrow temperature interval where the transition takes place. The
processing of the phase spectrum to obtain the dispersion curve suggests that
the velocity is almost independent of frequency in the interval 5 - 40 MHz.
PM AFM Transition

1.0
~
.,
-0
.
~0.5
<t:

o 10 20 30 40 50 60 70 o 10 20 30 40 50 60 70 o 10 20 30 40 50 60 70
Frequency (MHz) Frequency (MHz) Frequency (MHz)
(a) (b) (c)

Figure 5. 15 Spectrum of the backwall ultrasonic echoes for the PM phase, AFM phase
and for the transition. Three curves are shown for each phase to show the reproducibility
(de Camargo et al. ,2001b).

The physical process that causes the ultrasonic attenuation near the Neel
phase transition in chromium alloys is not established (Fawcett, 1988). The
use of ultrasonic spectral analysis offers some advantages as compared to the
methods that use single frequency. A single measurement with a wide band
transducer and spectral analysis allows one to get information for a continuous
frequency interval, and therefore, contains data which would require several
Spin-Density Waves and Charge-Density Waves in Cr Alloys 201

1.0..---------1""---

Figure 5. 16 Ultrasonic echoes spectrum as a function of the temperature (de Camargo


et al. 2001b).

experimental runs if performed with single frequency ultrasonic apparatus. The


single frequency technique is limited to the odd harmonics of the transducers
and requires replacement of the transducer with new coupling and other
adjustments that make comparison unreliable. The water immersion has the
advantage that does not cause any mechanical stress and, of course, the
disadvantage of a limited temperature range of operation. The spectrum
presented in Fig. 5.14 shows the wide band response from 5 IJHz to 40 MHz
where the expected decrease in amplitude for large frequencies is observed.
Considering the backwall and the entrance echoes, it is clear that there is a
general trend of decreasing amplitude as the frequency increases.
The ultrasonic attenuation is smaller in the paramagnetic phase than in the
antiferromagnetic phase, in agreement with single frequency measurements
(Castro et al., 1986) and shows weak frequency dependence. The larger
attenuation in the antiferromagneitc phase is likely to be due to the magnetic
domains (Castro et al., 1986). The energy loss due to magneto-elastic
coupling between the spin density wave and an oscillatory elastic strain on
chromium and on its antiferromagnetic alloys cannot be represented by a
simple interaction between static domains and the ultrasonic pulse. The aim of
this section in this chapter is to show some advantages of the spectral analysis
over the single frequency technique.
The ultrasonic spectral analysis is shown to be simpler and to provide
much richer results than the single frequency measurements usually used for the
202 A. J. A. de Oliveira and P. C. de Camargo

investigation of phase transition in solid materials like in the case of


Cr-O. 18 at. %. It is clearly shown that large energy loss at the magnetic phase
transition in Cr-O. 18 at. % Re is strongly frequency dependent, but has a weak
frequency dependence in the range 5 to 40 MHz for the paramagnetic and
for the antiferromagnetic. The 3-D representation amplitude-frequency-
temperature (Fig. 5. 16) provides a global view of the phenomena allowing
one to choose the best frequency and temperature to be used in order to
investigate the physical parameter of interest. The sound velocity is shown to
be almost frequency independent. The area under each ultrasonic pulse,
containing all frequencies of the wide band transducer and detecting system, is
a parameter that represents the energy loss and can be used to compare the
behavior of other physical quantities around the phase transition.

5.7 Bulk Chromium Properties Relevant for Thin


Film Devices

Surface magnetism of 3d transition metals is an interesting field of


investigation. The reduction of dimensionality induces new ordering on the
surface of otherwise paramagnetic or antiferromagnetic bulk crystals. In the
case of Cr the surface ferromagnetism on (001) direction is associated with
the high density of states that arises from localized surface-state bands near
the Fermi level (Fu and Freeman, 1986).
There are many reports in literatures showing the existence of ferromagnetic
surface layer in antiferromagnetic chromium (Ferguson, 1978; Matsuo and Nishida,
1980; Klebanoff et ai., 1985; Fitzsimmons et ai., 1994). Allan (1978, 1979,
1981) predicted that the reduced coordination number 4 at the Cr (001) surface
would result in surface magnetic order and the ferromagnetic surface phase
characterized by an exchange-split surface spin density of states with large localized
surface magnetic moments of about 2.8 Ps.
The surface ferromagnetism in Cr may be understood through Fig. 5. 17
(Klebanoff et al., 1985, Klebanoff and Shirley, 1986). The single-Q
antiferromagnetism in bcc structure corresponds to the polarization in alternate
(DOl) planes being antiparallel, so that the surface (DOl) plane contains
atoms having the same spin direction. Klebanoff et al. , the using formalism of
the Slater-Koster tight-binding scheme, including s, p, and d orbitals with
interactions up to second nearest neighbors calculate the magnetic moment on
surface of the Cr. The first layer has the estimated value of 3.00 Ps' and the
sixth layer with 0.85 Ps' very close to bulk value of the Cr, which is 0.62 Ps
As discussed above (Section 5. 5) many different works show the
Spin-Density Waves and Charge-Density Waves in Cr Alloys 203

Figure 5. 17 Magnetic moments of surface and near-surface atoms at the Cr ( 100 )


surface. The diameter of the circles represents the magnitude of the moment Ii: surface
layer iiI = 3. 00 liB; second layer, 1i2 = -1.56 liB' 1i3 = 1.00 liB' 1i4 = - O. 93 liB' 1i5 =
- 0.86 liB' 1i6 = 0.85 liB' Solid (open) circles indicate polarization parallel (antiparallel)
to (100) or (010) (Victoria and Falicov, 1985).

existence of single-Q state near the surface of Cr (Zabel, 1999; de Camargo


et al. , 2001 a; Mannix et al. , 2001). The single-Q state can be induced in
bulk chromium applying high magnetic field or stress. Magnetic field applied
along, for example (100) direction, above the Neel temperature, sample
cooled through TN and removed below TN' In this case the high anisotropy is
created, because the SOW is polarized along this direction, i. e., the Q
vector parallel to direction of applied magnetic field. In fact, all recent X-ray
results have shown Q perpendicular to the (00l) surface. We believed that the
origin of this anisotropy in the surface of Cr is the relationship with the surface
ferromagnetism, as describe above. As Cr is used as spacer layer in many
magnetic multilayers, the surface ferromagnetism present in Cr is very
important to understand the mechanism of scattering in interface. Fullerton
et al. (1996) observed through neutrons scattering the persistence of SOW in
multilayers of Fe/Cr (001) and its coupling through Fe layers. The Neel
temperature decreases with thickness of the Cr layer, vanishes to 0 K. As a
consequence, there is the suppression of the SOW state, for thickness less
than 42 A (Fullerton et aI., 1995).
An example of this effect was observed by Pflaum et al. (1999) through
magnetization as a function of magnetic field and ferromagnetic resonance.
The coercive field of a thin film of Fe (20 A) /Cr ( 10, 120 A) presents a larger
variation as a function of temperature. This effect is associated to spin-flip
transition. An increase of coercive field was observed for temperatures below
100 K, which can be related to a reorientation of the magnetic moments in Cr
from an out-of-plane to an in-plane alignment accompanied by an onset of
coupling at the Fe/Cr interface. The latter leads to a sensitivity of the
magnetic domains in the Fe layer on the magnetic order at the Fe/Cr
interface.
204 A. J. A. de Oliveira and P. C. de Camargo

5.8 Final Remarks

The interest in the magnetism of Cr and its antiferromagnetic alloys, since its
discovery in the 1930s, has stimulated many investigations. The fact that Cr is
an itinerant antiferromagnet with no direct orbital moment (L) contribution,
and its SOW being incommensurate with reciprocal lattice, together with the
COW/SW transform this subject into a very special case.
Many magnetic properties of Cr alloys show a strong dependence of
electron/atom ratio, because increasing (or decreasing) the electron
concentration improves the nesting of the Fermi surfaces responsible for the
electron-hole condensation that gives rise to the SOW. Alloying with metals of
the groups 7 and 8 increases the wave vector Q and the Neel temperature with
the exceptions of Fe and Co, while TN and wave vector decreases with
increasing concentration of group-4 and -5 metals.
The richness of magnetic phase diagram of Cr alloys was explored in
detail by Fawcett et al. (1994) showing that the magnetic properties of these
systems can be controlled with of the addition of impurities of transition
metals. Cr-based alloys can be prepared to be antiferromagnetic,
ferromagnetic and paramagnetic. More recently, the discovery of coupling of
the SOW in magnetic multi layers encourages the investigation of transport
properties in systems that exhibit the SOW and how the modulation of wave
vector modify the magneto transport properties. The review by Zabel (1999)
on chromium thin films explores the aspects of interlayer magnetic coupling and
also the magnetic Q domains; however, the effects of alloying in thin films still
lacks systematic research.
The recent observation of the Curie-Weiss behavior with the addition of
nonmagnetic impurities, especially V, represents the new challenge in the
determination of its origin that is likely to be due to the formation of local
moments connected to a local SOW. As far as the authors' knowledge goes,
there is no other report relating the appearance of local magnetic moments in
SOW alloys with the introduction of nonmagnetic impurities. The similar
systems where non-magnetic impurities give rise to local moments like Heusler
alloys or as in y-Fe and a-Mn, or borocarbides (Lynn et aI., 1997> do not
involve SOW.
Effects such as spin-glass phase in Cr-Mn alloys (Galkin et aI., 1995),
giant magnetoresistance in Cr-Fe-V and Cr-Fe-Mn alloys (Somsen et al.,
2000), thermal fluctuations in Cr-Fe-V and Cr-Co-V alloys, and exchange bias
behavior in Cr-Co-V alloys, are also related with the formation of local SOW
around of defects and impurities in Cr alloys.
The technology of magnetic thin films may take advantage of aspects like
Spin-Density Waves and Charge-Density Waves in Cr Alloys 205

spin glass characteristics and local moments formation in bulk materials, even
though this is certainly a non-trivial task.
X-ray and neutrons are complementary tolls that have to be used for a full
understanding of the distribution of Q domains and S domains relevant for the
their determination, which can be helpfull on understanding spin controlling
devices. A carefull mapping of the distribution of Q and S domains and their
dependence on sample geometry, defects and applied magnetic field can be of
great importance in the understanding of main features of the spin behavior in
metals and its interfaces.
The persistence of the effects of spin-fluctuations in paramagnetic phase
of Cr and of SOW alloys, up to temperatures well above the Neel transition is
intriguing and may be related to the fluctuations observed in high-temperatures
superconductors.
Finally, the spin-density waves and charge-density waves in Cr alloys
present a complex and interesting subject in magnetism. The separation of the
contribution due to an intrinsic COW, from that of an SW may require crystals
of exceptionally good crystalline quality, a variety of alloys and a combined X-
ray and neutrons experiment. The possibility of understanding the phenomenon
of the SOW and its connections to the ordered magnetic systems is a most
important point for the research in these alloys systems.

References
Aidun R. , S. Arajs and C. A. Moyer. Phys. Stat. Solidi (b) 128: 133(1985)
Anderson, P. W. Advances in Physics 46: 3(1997)
Ando, M. and S. Hosoya. J. Appl. Phys. 49: 6045(1978)
Alberts, H. L. and J. A. J. Lourens. Phys. Rev. B 29: 5279(1984)
Alberts, H. L. and J. A. J. Lourens. Journal de Physique C-C8: 215 ( 1988)
Alberts, H. L. and A. H. Boshoff. J. Magn. Magn. Mat. 104-107: 2031
( 1992a)
Alberts, H. L. and J. A. J. Lourens. J. Phys.: Condens. Mattter 4: 3835
(1992b)
Allan, G. Surf. Sci. 74: 79(1978a)
Allan, G. Surf. Sci. Rep. 1: 12l(1978b)
Allan, G. Phys. Rev. B 19: 4774(1979)
Allen, J. W. and C. Y. Young. In: C. G. Graham Jr. , G. H. Lander and J.
J. Rhyne, eds. Magn. Magn. Mat. AlP Cont. Proc., AlP NY, 24: 410
(1975)
Allen, J. W. and C. Y. Young. Phys. Rev. B 16: 1103(1977)
Arajs, S. , C. A. Moyer, J. R. Kelly and K. V. Rao. Phys. Status Solid B
101: 2747(1975)
Arajs, S., G. Kote, C. A. Moyer, J. R. Kelly, K. V. Rao and E. E.
Anderson. Phys. Status Solidi B 74: K23( 1976)
Arajs, S., C. A. Moyer and O. Abukay. Phys. Status Solid BIOI: 63(1980)
206 A. J. A. de Oliveira and P. C. de Camargo

Baibich, M. N., J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P.


Etienne, G. Cruzet, A. Friederich and J. Chazelas. Phys. Rev. Lett. 61:
2472(1988)
Baker, A. S. and J. A. Ditzenberger. Phys. Rev. B 1: 4378(1970)
Baran, A. , H. L. Alberts, A. M. Strydom and P. de V. du Plessis. Phys.
Rev. B 45: 10,473(1992)
Baruchel, J. , P. C. de Camargo, H. Klein, I. Mazzaro, J. Nogues and A.
J. A. de Oliveira. J. Phys. 0: Appl. Phys. 34: Al14(2001)
Bender, D. and J. Muller. Phys. Kondens. Matter 10: 342 ( 1970)
Benediktsson, G., H. U. Astrom, K. V. Rao. J. Phys. F 5: 1966(1975)
Benediktsson, G. , L. Hedman, H. U. Astrom, K. V. Rao. J. Phys. F 5:
1439 (1982)
B6deker, P. , A. Hucht, A. Schreyer, J. Borchers, F. GUthoff and H. Zabel.
Phys. Rev. Lett. 81: 914(1998)
B6deker, P., A. SchreyerandH. Zabel. Phys. Rev. B59: 9408(1999)
Bosshoff, A. H., H. L. Alberts P. V. du Plessis and A. M. Venter. J. of
Phys. Condo Matter 5: 5353( 1993)
Burke, S. K. and B. D. Rainford. J. Phys. F 8: L239(1978)
Burke, S. K. and B. D. Rainford. J. Phys. F 13: 441(1983)
Bridgman, P. W. Proc. Am. Acad. Arts. Sci, 68: 27(1932)
Brunei, M and F. de Bergevin. In: G. S. Brown and D. E. Moncton, eds.
Magnetic Scattering, of the Handbook of Synchrotron Radiation. North
Holland, Vol 03 in the chapter 14, (1991)
Brux, U. , T. Schneider, M. Acet and E. F. Wassermann. Phys. Rev. B 52:
3042(1995)
Cable, J. W., S. A. Werner, G. P. Felcher and N. Wakabatashi. Phys.
Rev. Lett. 49: 829 ( 1982)
Cable, J. W., S. A. Werner, G. P. Felcher and N. Wakabatashi. Phys.
Rev. B 29: 1268(1984)
Castro, E. P., P. C. de Camargo and F. R. Brotzen. Sol id State
Communications 57: 37 ( 1986)
Chiu, C. H., M. H. Jericho, R. H. March. Can. J. Phys. 49: 3010 (1971)
Creveling, L. Jr. and H. L. Luo. Phys. Rev. 176: 14(1968)
Creveling, L. Jr. and H. L. Luo. Phys. Lett. A 28: 72 ( 1969)
de Bergevin, F. and M. BruneI. Phys. Lett. A 39: 141( 1970)
de Bergevin, F. and M. BruneI. Acta Cryst. A 37: 314 ( 1981)
de Camargo, P. C., E. P. Castro and E. Fawcett. J. Phys F 18: L209
(1988)
de Camargo, P. C. , E. Fawcett and J. M. Perz. J. Appl. Phys. 67: 5265
(1990)
de Camargo, P. C., A. J. A. de Oliveira, C. Giles, F. Yokaichiya, C.
Vettier. Mater. Sci. Forum 302-303: 32(1999)
de Camargo, P. C., I. Mazzaro, C. Giles, F. Yokaichiya, A. J. A. de
Oliveira, H. Klein and J. Baruchel. J. Magn. Magn. Mat. 233: 65(2001a)
Spin-Density Waves and Charge-Density Waves in Cr Alloys 207

de Camargo, P. C. , S. E. Kruger and M. A. Rebello. Scripta Materialia 44:


2373(2001b)
de Faria, C. C. , A. J. A. de Oliveira, F. M. A. Moreira and W. A. Ortiz.
IEEE Trans. on Magn. 31-36: 3403(1995)
de Oliveira, A. J. A. , O. F. de Lima, W. A. Ortiz and P. C. de Camargo.
Sol id State Comm. 96: 383 ( 1995)
de Oliveira, A. J. A., W. A. Ortiz, P. C. de Camargo and V. Yu. Galkin.
J. Magn. Magn. Mat. 152: 86(1996a)
de Oliveira, A. J. A. , O. F. de Lima, W. A. Ortiz, P. C. de Camargo and
E. Fawcett. Journal of Physics: Condensed Mat. 8: L403(1996b)
de Oliveira, A. J. A. , W. A. Ortiz, O. F. de Lima and P. C. de Camargo.
Journal Appl. Physics 81: 4209 (1997)
de Oliveira, A. J. A., W. A. Ortiz and P. C. de Camargo. Materials
Science Forum 302-303: 344(1999)
de Oliveira A. J. A, W. A. Ortiz and P. C. de Camargo. Journal of Magn.
and Magn. Mat. 226-230: 1086(2001)
deOliveira, L. M. W.A.Ortiz and A. Oliveira. J. Appl. Phys. 93(10): 7154
(2004)
Eisebitt, S. , T. Boske, J. R. Rubensson, E. Eberghardt. Phys. Rev. B 60:
10,170(1993)
Endoh, Y. , J. Mizuki and Y. Ishikawa. J. Phys. Soc. Jpn. 51: 2826( 1982)
Fawcett, E. Rev. of Mod. Phys. 60: 253(1988)
Fawcett, E., D. R. Noakes. Int. J. Mod. Phys. B 7: 624(1993)
Fawcett, E., R. B. Robert, R. Day and G. K. White. Europhys. Lett. 1:
473(1986)
Fawcett, E., H. L. Alberts, V. Yu. Galkin, D. R. Noakes and J. V.
Yakhmi. Rev. of Mod. Phys. 66: 25(1994)
Fawcett, E., V. Yu. Galkin and W. A. Ortiz. Magn. Magn. Mat. 198-199:
425(1999)
Fedders, P. A. and P. C. Martin. Phys. Rev. 143: 245(1966)
Fishman, R. S. and S. H. Liu. Phys. Rev. B(1992)
Fishman, R. S. J. Phys.: Condens. Matter 13: R235(2001)
Ferguson, P. E. J. Appl. Phys 49: 2203(1978)
Fitzsimmons, M. R. , J. A. Eastman, R. A. Robinson, A. C. Lawson, J. D.
Thompson, R. Movshovich and J. Satti. Phys. Rev. B 48: 8245(1993)
Fitzsimmons, M. R., J. A. Eastman, R. B. Von Dreele and L. J. Thompson.
Phys. Rev. B 50: 5600(1994)
Friedel, J. and L. E. Hedman. J. Phys. (Paris) 39: 1225(1978)
Fu, C. L. and A. J. Freeman. Phys. Rev. B 33: 1755(1986)
Fullerton, E. E., K. T. Riggs, C. H. Sowers, S. D. Bader and A. Berger.
Phys. Rev. Lett. 75: 330(1995)
Fullerton, E. E., S. D. Bader, J. L. Robertson. Phys. Rev. Lett. 77: 1382
(1996)
Gabovich, A. M., A. I. Voitenko, J. F. Annett, M. Ausloos. Supercond.
208 A. J. A. de Oliveira and P. C. de Camargo

Sci. Tech. 14: R1(200 1)


Galkin, V. Yu., P. C. de Camargo, N. Ali, J. Schaf and E. Fawcett. J.
Phys. Condens. Mater 7: L649( 1995)
Galkin, V. Yu. P. C. de Camargo, N. Ali and E. Fawcett. J. Phys.
Condens. Mater 8: 7925(1996)
Galkin, V. Yu et al. J. Phys.: Condo Mat. 9: L577(1997)
Galkin, V. Yu., W. A. Ortiz and E. Fawcett. Journal Applied Physics 83:
7384(1998)
Galkin, V. Yu., W. A. Ortiz and E. Fawcett. J. Appl. Phys. 87: 6543
(2000)
Galkin, V. Yu., W. A. Ortiz, N. Ali and E. Fawcett. Magn. Magn. Mat.
226-230: (2001)
Geerken, B. M., R. Grissen, G. Benediktsson, H. U. Astrom and C. van
Dijk. J. Phys. F 12: 1603(1982)
Gibbs, D. G. Grubel, D. R. Harshmann, E. D. Isaacs, D. B. Mcwhan, D.
Mills and C. Vettier. Phys. Rev. B 43: 5663(1991)
Grier, B. H., G. Shirane and S. A. Werner. Phys. Rev. B31: 2882(1985)
Grunberg, P., R. Schreiber, Y, Pang, M. B. Brodosky and H. Sowers.
Phys. Rev. Lett. 57: 2442(1986)
Gruner, G. Rev. of Mod. Phys. 66: 1( 1994)
Harders, T. M. and P. Wells. J. Phys. F 13: 1017(1973)
Hattox, T. M., J. B. Conclin, J. C. Slater and S. B. Trichi. J. Phys.
Chem. Solids 34: 1627(1973)
Hedgcock, F. T., J. O. Strom-Olsen, D. F. Wilford. J. Phys. F: metal
Phys. 7: 855(1977)
Hedman, L., K. Svenson, K. V. Rao, S. Arajs and H. U. Astrom. J. Low
Temp. Phys 14: 545(1974)
Hill, P. , N. Ali, A. J. A. de Oliveira, W. A. Ortiz, P. C. de Camargo and
E. Fawcett. J. Phys. : Condens Matter 6: 1761(1994)
Hill, J. P., G. Helgensen, D. Gibbs. Phys. Rev. B 51: 10,336(1995)
Hill, J. P. and D. F. McMorrow. Acta Crystallogr. A 52: 236(1996)
Hirano, M. and Y. Tsunoda. Phys. Rev. B 59: 13,835(1999)
Iida, S., S. Kawarazaki and N. Kunitomi. J. Phys. Soc. Jpn. 50: 3612
(1981)
Ishikawa, Y., R. Tounier and J. Filippi. J. Phys. Chem. Solids 26: 1727
(1965)
Kelly, J. R. C. A. Moyer and S. Arajs. Phys. Rev. B 20: 1099(1979)
Klebanoff, L. E., R. H. Victoria, L. M. Falicov and D. A. Shirley. Phys.
Rev. B 32: 1997(1985)
Kleiber, M. , M. Bode, R. Ravlic and R. Wiesendanger. Phys. Rev. Lett.
85: 4606(2000)
Koehler, W. C., R. M. Moon, A. L. Trego and A. R. Macintosh. Phys.
Rev. 151: 405(1966)
Kondorskiy, E. I., T. I. Kostina, V. P. Medevedchikov and Yu. A.
Spin-Density Waves and Charge-Density Waves in Cr Alloys 209

Kuskova. Phys. Met. Metall. 48: 27(1980


Langridge, S., G. H. Lander, N. Bernhoeft, A. Stunault, C. Vettier, G.
Grubel, S. Sutter, F. de Bergevein, W. J. Nuttall, W. G. Stirling, K.
Mattenburger and O. Vogt. Phys. Rev. B 55: 6392( 1997)
Lebech, B. and K. Mikke. J. Phys. Chem. Solids 33: 1651(1972)
Lomer, W. M. Proc. Phys. Soc. London 80: 489(1962)
Lynn, J. W, A. Skanthakumar, Q. Huang, S. K. Sinha, Z. Hossain, L. C.
Gupta, R. Nagarajan and C. Godart. Phys. Rev. B 55: 6584(1997)
Maki, S. and K. Adachi. J. Phys Soc. Japan 46: 1131( 1979)
Mannix, D., P. C. de Camargo, C. Giles, A. J. A. de Oliveira, F.
Yokaichiya and C. Vettier. Eur. Phys. J. B 20: 19(2000
Mason, T. E., G. Aeppli and H. A. Mook. Phys. Rev. Lett. 68: 1414
(1992)
Matsuo, S. and I. Nishida. J. Phys. Soc. Jpn. 49: 1005( 1980)
Michel, R. P. et al. Phys. Rev. B 44: 7413(1990
Michel, R. P. et al. Phys. Rev. B 47: 3442( 1993)
Muheim, S. and J. Muller. Phys. Kondens. Mater 2: 377(1964)
Muir, W. C., E. Fawcett and J. M. Perz. Phys. Rev. Lett. 59: 335(1987a)
Muir, W. C. , J. M. Perz and E. Fawcett. J. Phys. F: Met. Phys. 17: 2431
(1987b)
Mydosh, J. A. Spin Glasses: An Experimental Introduction. Taylor &
Francis, London( 1993)
Nait-Laziz, H., C. Demagrat and A. Mobram. J. Magn. Magn. Mater. 121:
123(1993)
Neel, L. C. R. Acad. Sci. 203: 304(1936)
Neubeck, W. PhD. Thesis, Universite Joseph Fourier, Grenoble(2000)
Neubeck, W., C. Vettier, F. de Bervegin, F. Yakhou, D. Mannix, O.
Bengone, M. Alouani, A. Barbier. 2001(to be published)
Nishihara, Y. , Y. Yamauchi, M. Tokumoto, K. Takeda and F. Fukamichi.
Phys. Rev. B 34: 3446(1986)
Noakes, D. R., T. M. Holden and E. Fawcett. J. Appl. Phys. 67: 5262
( 1990a)
Noakes, D. R., T. M. Holden, E. Fawcett and P. C. de Camargo. Phys.
Rev. Lett. 65: 369(1990b)
Nogues, J. and I. K. Schuller. J. Magn. Magn. Mater 192: 203(1999)
Overhauser, A. W. Phys. Rev. 128: 1437(1962)
Overhauser, A. W. Phys. Rev. Lett. 4: 226(1960)
Pflaum J. , J. Pelzl, Z. Frait, P. Sturc, M. Marysko, P. B6deker, K. Theis-
Br6hl and H. Zabel. J. Magn. Magn. Mat. 198-199: 453(1999)
Rau, C., G. Xing, C. Liu and M. Robert. Phys Lett. A 135: 227(1989)
Renninger, M. Z. Phys. 106: 141(1937)
Rice, T. M. Phys. Rev. B 2: 3619(1970)
Sato, H. and K. Maki. Int. J. Magn.6: 193(1974)
Sidek, H. A. A., M. Cankurtaran, G. A. Saunders, P. J. Ford and H. L.
210 A. J. A. de Oliveira and P. C. de Camargo

Alberts. Phys. Lett. A 172: 387(1993)


Shibatani, A. , K. Motizuki and T. Nagamiya. Phys. Rev. 177: 984( 1969)
Shull, C. G. and M. K. Wilkinson. Rev. Mod. Phys. 25: 100(1953)
Somsen, Ch., M. Acet, G. Nepecks and E. F. Wassermann. J. Magn.
Magn. Mater. 208: 191(2000)
Squires, G. L. Introduction to the Theory of Thermal Neutron Scattering.
Cambridge University, Cambridge, England(1978)
Stassis, C. , G. R. Kleine and S. K. Sina. Phys. Rev. B 11: 2171(1975)
Steinitz, M.a. et al. Phys. Rev. Lett. 23(17): 979(1969)
Stockert, O. , S. M. Hayden, B. Fak and G. Aeppli. Physica B 281: 701-
702(2000)
Strempfer, J., Th. Bruckel, G. J. Mcintyre, F. Tasset, Th. Zeiske, K.
Bruger and W. Prandl. Physica B 267: 56 (1999)
Strempfer, J., Th. Bruckel, W. Caliebe, G. J. Mcintyre, F. Tasset, Th.
Zeiske, K. Bruger, W. Prandl and J. R. Schneider. Eur Phys. J. B 14: 63
(2000)
Street, R, B. C. Munday, B. Window and J. R. Wi II iams and J. Appl. Phys.
39: 1050(1968)
Strom-Olsen, J. 0., D. F. Wilford, S. K. Burke and B. D. Rainford. J.
Phys. F 9: L95(1979)
Stunault, A., F. de Bergevin, D. Wermeille, C. Vettier, T. Brukel, N.
Bernhoeft, G. S. Mcintyre, J. Y. Henry. Phys. Rev. B 47: 10,170(1999)
Suzuki, T. J. Phys. Soc. Jpn. 21: 442(1966)
Trego, A. L. and A. R. Mackintosh. Phys. Rev. 166: 495(1968)
Tugushev, V. V. In: W. Hanke and Yu. V. Kopaev, eds. Modulated and
Localized Structures of the Spin-Density Wave in Itinerant Antiferromagnets
in Electronic Phase Transitions. Elsevier, New York/Amsterdam, p. 237
(1992)
Vaillant, F. Acta Crystallogr. A 33: 967(1977)
Walker, M. B. Phys. Rev. Lett. 44: 1261(1980)
Werner, S. A. Comments Condens. Matter. Phys. 15: 55 ( 1990)
Werner S. A., A. Arrot, H. Kendrick. J. Appl. Phys. 37: 1260(1966)
Williams, I. S., E. S. R. Gopal and R. Street. J. Phys: Metal Phys. F 3:
431(1979)
Zabel, H. J. Phys.: Condens.Mater 11: 9303(1999)
6 New Magnetic Recording Media

D. J. Sellmyer, H. Zeng, M. Van, S. Sun and Y. Liu

6. 1 Introduction

Magnetic recording media can take the form of tape, floppy disk, or hard
(rigid) disk. The signals recorded could be analog or digital. In computer data
storage applications, all the data are stored in digital form. Tapes and floppy
disks are used for permanent storage and hard disks are used for on-line
storage. The principles of these recording methods are well described in the
magnetic storage handbook and other reference books (Mee and Daniel,
1996). This chapter will focus on the recording media for hard disk drives,
which have experienced the most significant development in recent years and
represent the most technical challenges.
Brief history of magnetic thin film development: Magnetic thin films have
been used as the recording medium in hard-disk drives for about fifteen years.
The progress of areal density for thin-film media is shown in Fig. 6. 1. There
have been three distinct periods in the development of thin-film materials
associated with the reading head available. As the signal retrieval of the
inductive head is weak, the first period was focused on depositing films with
the proper coercivity (He) and remanence-thickness product (M, t) to
enhance read-back signal. The second stage of thin-film media developments
emerged with the introduction of the magnetoresistive (MR) head technology
in 1990. Because of the great read-back sensitivity of the MR head, signal-to-
2
10 -------------------------------------
~ 10 1
:;; 10
:0
8 Wl
.f' w 2
</>

~ 10- 3
] 10-4
.:;: 10- 5
10- 6 ' - - _ - ' - - _ - - ' - _ - - ' - _ - - - ' -_ _' - - _ - ' -
1950 1960 1970 1980 1990 2000 2010
Year of product introduction
Figure 6. 1 Progress of areal density in magnetic recording.
212 D.J.Sellmyer et al.

noise ratio, instead of signal amplitude, became the limiting factor to drive
performance. The research effort in this period focused on media
microstructure, namely, the grain-size refinement. The grains also need to be
isolated by either chemical or physical means to reduce intergranular coupling.
The recent introduction of the giant magnetoresisitive (GMR) head benchmarks
the third era of thin-film media design. In this period research focused on the
refinement of grain size and grain-size distribution. Superparamagnetic effects
concomitant with small grains become the ultimate limit for further increasing
areal densities.
Two critical factors for ultimate high-density thin film design are the
signal-to-noise ratio and thermal stability of the film. These two factors can be
related to grain size, grain-size distribution, texture, magnetization, grain
isolation, and film smoothness. The noise sources include reading head noise
and medium noise. The signal to medium noise ratio is given by

SIN = O.31PW
2
50 BW (6.1)
a sO g
where PW50 is the pulse width of the read back signal, B is the spacing
between two bits, W is the track width and s is the cross track correlation
width (Bertram et al. , 1998). s is the measure for magnetization fluctuation in
the direction across the recording track. a is a parameter determining the
magnetic transition length and is defined by grain size and the transition
parameter. Og is a parameter affected by the grain size and texture of the
medium. a is given by

(6.2)

where 0 is the grain size. awe is the transition parameter (Zhou and Bertram,
1999) and is affected by the squareness of the magnetic loop of the medium,
magnetic transition shape and the head field gradient.
The grain-size distribution also plays a critical role in affecting the noise
as described by (Zhou et aI., 1999)

(6.3)

where a is the width of grain size distribution.


Thermal stability of future media design is becoming increasingly
important. The thermal-activation model and deduced the coercivity used as a
function of time (Sharrock, 1994; Lu and Charap, 1994),

He(t) = H o {1 - [~: ~ln(At) r}, (6.4)

where k B' is the Boltzmann constant, T is the temperature, K u is the


magnetic anisotropy of the medium, and V is the grain volume, respectively;
A is a time-independent constant and n varies from 1/2 to 2/3 depending on
New Magnetic Recording Media 213

the orientation distribution of grains. To satisfy the thermal stability


requirement, the condition should be sustained for a small grain medium.

Ku V > 60 (6.5)
ksT
where V represents the smallest grain size.
Current approaches for new recording media development take three
routes: The first approach uses the sputtering technique, which has been well
established. There are two challenges in this approach. One is how to reduce
the grain size and the other is how to sustain signal-to-noise ratio. The
limitation of this approach is that the grain size reduction has almost
approached its limit. The second approach uses patterned media in which one
grain acts as a recording bit. This approach has the advantage to increase the
recording density by a factor of 500 - 1000. The technical difficulties are how
to fabricate large area patterned media and how to read/write signals on such
media. The third approach is to use patterned perpendicular recording media.
As the magnetization along the z direction is perpendicular to the disk surface,
the recording unit has no restriction in the z direction. This allows larger
volume and the shape anisotropy to stabilize the magnetic structure. There are
also technical difficulties in perpendicular recording media. This chapter will
review recent media research thrust areas by sputtering, chemical synthesis
and self-assembly.

6. 2 New Media Development by Sputtering

6. 2. 1 Introduction
The first generation of magnetic media for hard disks used particulate media
consisting of y-Fez 03 particles dispersed in polymeric binders. The technology
to fabricate such media is that the coating was "spun" onto the substrates and
magnetic field was applied circumferentially to align magnetic particles before
the solvents had completely evaporated. For two purposes, aluminum oxide
particles with diameters larger than the coating thickness were added into the
magnetic particles: to keep the mechanical durability, and to provide
separation between the head and the polymeric binder. Because the particle
volume fraction needs to be below about 30 % in order to maintain adequate
mechanical properties of the composite coating, thicker coatings are required
to obtain sufficient read-back amplitude. The limitation of this media is that
thinner coatings with good uniformity are very difficult to obtain and the non-
magnetic aluminum oxide particles degraded the signal-to-noise ration of the
214 D.J.Sellmyer et al.

system.
Today most magnetic recording media for hard-disk drives are thin films,
which are deposited by sputtering. The advantages of the sputtered film media
over particle-coated media are their superior magnetic properties with much
thinner film thickness as well as a smoother surface. A smoother surface
allows lower flying heights that reduce spacing between the head and the
medium to increase areal density. In addition, by modifying the sputtering
parameters, the microstructure can be controlled and magnetic properties can
be tailored to satisfy desired recording requirements. Other advantages of
sputtered thin films are higher magnetization, higher anisotropy, and
adjustable coercivity. Higher magnetization achieved by magnetic thin films
allows the use of a thinner recording layer with sufficient read-back signal.
Higher anisotropy favors the enhancement of thermal stability that is of
importance to raise areal density as well. Higher and tunable coercivities
obtained by adjusting the film composition allow the reduction of the transition
parameter for higher areal density. Figure 6.2 shows the basic structure of a
thin-film medium, which consists of a substrate, an underlayer, a magnetic
layer, an overcoat, and a lubricant. The AI-Mg alloy with a thick plated
amorphous NiP layer is used as the substrate. The hard NiP layer allows easy
polishing and provides a surface for resistance to mechanical damage. The
underlayer develops a necessary texture, which will control the grain size and
align the c-axis of the magnetic layer with hexagonal structure. Development
of new media by sputtering techniques has been focused on the following issues:
CD reduction of grain size, (2) uniform grain-size distribution, @ isolation of
magnetic grains, and @ synthesis of alloys with high magnetic anisotropy.

Lubricant
Overcoat

Figure 6.2 Basic structure of a thin film media.

New underlayer research: Cr and Cr-alloys (CrV, CrMo, CrW, etc.)


with body centered cubic (bcc) structure are used as underlayers to promote
the epitaxtial growth of magnetic layer. An additional nucleation or seed layer
is deposited before the Cr-alloy underlayer to control grain size and orientation
because it is more difficult to nucleate a necessary Cr texture on glass or glass-
ceramic surface. A lot of seed layers have been explored, such as NiAI (Lee
et aI., 1994), MgO (Lee et aI., 1995), Ta (Kataoka et aI., 1995) and
New Magnetic Recording Media 215

AIN (Mirzamaani and Doerner, 1996).

6.2.2 Hexagonal Co-Alloys


Hexagonal close packed (hcp) Co-alloys are commercially used magnetic
layers today. Most of the media materials used commercially today are ternary
and quaternary hexagonal Co-alloys. They are CoCrTa, CoCrPt, CoNiPt,
CoTaNi, CoCrPtTa, and CoPtCrX (X= Ni, B, or Si). Some investigations of
CoCrPtTaNb were also reported.
Reducing media noise is a key issue in magnetic recording. In thin film
media the noise source is localized at the transition zone that is of the so-called
"zig-zag" pattern. Transition noise strongly depends on the grain size and
whether the grains are exchange-coupled (Bertram and Zhu, 1992).
Recording parameters such as transition pulse width and bit aspect ratio are
another noise source (Arnoldussen et aI., 1999).
Reducing interactions between the magnetic grains will help to reduce the
media noise. The methods to reduce interactions between grains are
physically isolating or chemically segregating the grains.
Effect of sputtering parameters: In sputtering the common method is to
adjust deposition parameters or add other elements to Co-alloy. For example,
using high gas pressure to sputter Co-alloy can create voided grain
boundaries, which will help to isolate the grains physically (Yogi et al. ,
1990).
Decoupling of grains by segregation of non-magnetic element at grain
boundaries: Enhancing Cr concentration will produce more Cr in Co-alloy
grains boundaries, which will help to segregate grains chemically (Doerner et
aI., 1993). Various analytical techniques have been used to analyze the
segregation of Cr in Co-alloy (Maeda et aI., 1985; HoNo et aI., 1992) and
very clear images have been observed (Wittig et aI., 1998). High Cr in Co-
alloy forms non-magnetic grain boundaries, which reduce the exchange
coupling between neighboring Co-alloy grains. This reduces the transition
noise. Another role of Cr in Co-alloy is to increase the corrosion resistance as
a well-known chemical stability enhancer.
Some elements and oxides have also been added to Co to improve signal-
to-noise ratio and for other purposes. These include Ta, B, Nb, P, Pt, W,
Si, Ir, Sm, and CoO. Although Ta has not been observed to segregate, it is
claimed that segregation of Cr was enhanced with the addition of Ta, based on
the appearance of more amorphous-like grain boundaries in Ta containing
alloys (Nakai et al., 1994). The role of B in reducing noise is through grain
boundary segregation (Weller and Doerner et al. , 2000) as well as reduced
grain size (Kubota et aI., 1998; McKinlay, 1999). The addition of Pt to Co-
alloy is different from most of the elements, which decrease noise and the
magnetocrystalline anisotropy.
216 D.J.Sellmyer et al.

Magnetocrystalline anisotropy enhancement: Pt was added to Co-alloy to


increase magnetocrystalline anisotropy, which is generally required to provide high
coercivity and thermal stability in the Co-alloys (Ishikawa and Sinclair, 1996; lnaba
et aI., 1999).

6. 2. 3 Ll o Alloys

The alloys of the L 10 type have a "superlattice" structure of CuAu I-type. It has
alternating monolayers of different atoms. FePt and CoPt binary alloys are
suitable as recording media because of their higher anisotropy
(-10 7 erg/cm3 ) . This higher anisotropy comes from the significant spin-orbit
interaction that was produced by symmetry breaking of the interfaces in a
superlattice system. Figure 6. 3 shows the crystal structure of FePt and CoPt.
The FePt and CoPt system form both disordered fcc and ordered-faced-
centered tetragonal fct structures near the equi-atomic composition. It is
known that hard magnetic behavior of these systems is related to the crystal
phase transition. The as-deposited film has a disordered fcc structure which
transforms into an ordered fct (L 10 ) structure after annealing at high
temperature. CoPt grains in the range of 100 - 300 nm in diameter were
obtained by dc-magnetron sputtering CoPt film and, subsequently, annealing in
the temperature range 550 - 800 C. Films exhibit coercivities about 30 kOe
(Liou et al. , 1996) and 37 kOe (Liou et al. , 1996). A coercivity very close to
the theoretical maximum of 48 kOe was found in 12 - 15 nm thick, epitaxial FePt
(00l) films, grown on MgO substrate. Coercivities between 2 and 7 kOe with grain
size less than 15 nm have been obtained in sputtered Fe/Pt mulitlayered films
annealed between 200 and 400 "c (Luo et aI., 1996).

OPt

Fe or Co

Figure 6. 3 Crystal structure of the FePt or CoPta .

6.2.4 Rare Earth-Transition Metal Compounds


Rare earth-transition metal (RE-TM) compounds have anisotropies in the
10 8 erg/cm 3 range. The largest known anisotropy constant K 1 belongs to the
New Magnetic Recording Media 217

hexagonal compound SmCos (K 1 =2 x 108 erg/cm 3 ) , which is about 20 times


larger than hcp Co and 50 - 100 times larger that representative CoCrPt alloy.
This large anisotropy would allow grain size down to very small diameter
(a few nm) and still retain the necessary thermal stability. It is natural that
much larger areal densities, even in the Tb/inz regime or above, would be
possible if one only considered thermal stability reasons. The problems for
fabricating these films with hard magnetic properties are unfavorably high
processing temperatures during either growth or post annealing, which are
unacceptable for mass production in industry. Usually, as-deposited RE-TM
films are amorphous and do not have large crystalline anisotropy when they are
sputtered below 600 'C. Some work has been reported to try solving this
problem (Velu and Lambeth., 1991; Liu et aI., 1994; Shan et aI., 1996).
Coercivities in the range of 2 - 4 kOe were obtained when CoSm films were
deposited on Cr (110) underlayer at room temperature. The microstructure of
the CoSm layer is composed of nanocrystallites of close-packed structure with
about 5nm diameter distributed in an amorphous matrix, and volume fraction of
nanocrystallites varies depending on the deposition pressure. CoPr films
deposited on Cr underlayer at room temperature have a mostly amorphous
CoPr Layer with the coercivity about 300 - 400 Oe and a nanostructure similar
to CoSm films. However, after annealing at 400C for 20 min the CoPr layer is
nearly 100 % crystall ine with a grain size of about 10 nm as revealed by the
high resolution TEM images. Coercivities were enhanced to the range of
2 - 8 kOe depending on the film thickness and deposition conditions (Malhotra
et aI., 1996).

6. 2. 5 Nanocomposite Films
6.2.5. 1 hcp Structured Films
CoCrPtM alloys (M = Ta, Nb, B, C, etc.) are used widely as hard disk
media today, and are still important candidates for future extremely high
density recording media. In order to reduce the interactions between magnetic
grains, CoCrPtM alloys embedded in non-magnetic matrix may improve the
magnetic isolation and reduce the grain size. CoCrPt!SiOz granular films,
which have a multilayered structure with SiOz of 4% -14% volume fraction as
deposited, have achieved a coercivity as high as 5. 6 kOe and a grain size of
10- 12 nm after annealing at 580 'C for 10 min (Xu et al. , 2000). CoCrPt/C
multi layers with 14 % - 17 % volume fraction of C, annealed at 530C for 30
min, obtained coercivity as high as 4. 6 kOe (Xu et al., 2000). All these
granular films have low Pt content ( ....... 12 at. %) and have the Co ( 110) /Cr
(002) texture.
6.2.5.2 fct Structured Films
L 10 phases or face-centered-tetragonal structure of near equi-atomic
218 D.J.Sellmyer et al.

composition CoPt and FePt embedded in non-magnetic matrix(C, Si02 , Ag,


A1 2 0 3 , B2 0 3 etc.) have been studied recently. Most of these films are
deposited in a multilayered film and subsequently annealed to obtain a
nanocomposite structure. Nanocomposite FePt: Si02 films have the coercivity
2 - 8 kOe and grain size 10 nm or less (Luo and Sellmyer, 1999). Good grain
isolation and narrow distribution were also achieved. In order to form L 10
phase, the as-deposited film needs to be annealed at temperature from 450 to
650 'c. The coercivity and grain size are highly dependent on the annealing
temperature and Si02 concentration. Annealed at temperatures above 600 "C ,
nanocomposite CoPt: C films (Yu et aI., 1999) with coercivities 3 - 12 kOe
have been obtained. Because neither Co and C nor Pt and C are miscible, Cis
an ideal isolation material between neighboring CoPt grains. The properties of
CoPt: C films are tailored toward magnetic recording applications and
recording linear density up to 10 kfc/mm were obtained. The c axes of the
FePt grains can be controlled both in or out of the film plane in nanocomposite
FePt: B2 0 3 films annealed at temperature 550'C (Luo et al., 2000). The
development of (001) texture depends strongly on the total film thickness,
initial B2 0 3 layer thickness and Fe concentration. The magnetic properties are
stable and coercivities can be tailored from 4 to 12 kOe with grain size down to
4 nm. For annealed CoPt: B2 0 3 composite films, the c-axis orientation of the
CoPt grains can also be controlled by adjusting the Co content (Yan et al. ,
2001 ). Nanocrytalline, physically separated FePt L 10 particles with
coercivities 4.4 kOe have been obtained by depositing extremely thin layers of
Pt (1. 5 nm) and Fe (1 nm), which were in situ coated with Ab 0 3 and
subsequently annealed (Bian et al. , 2000).

6. 2. 6 AFC Media
Antiferromagnetically coupled (AFC) media developed recently are expected
to permit hard-disk drives to store 100 Gb/in. and beyond very soon (Fullerton
et al. , 2000). AFC media consist of a sputtered multilayer structure in which
two magnetic Co-alloy layers are separated by a nonmagnetic ruthenium layer
three atoms thick. This thickness of the ruthenium causes the magnetization in
each of the magnetic layers to be coupled anti-parallel to each other.
Figure 6.4 shows the schematic representation of traditional media and AFC
media. For this structure, the effective areal moment density (M, t) of the
composite structure is the difference between the two ferromagnetic layers.
This allows M, t to be independent of the physical thickness of the media, and
permits the use of thicker and more thermally stable magnetic layers for a
given M, t value (Fullerton et al. , 2000). For future extremely high density
recording, high magnetocrystalline anisotropy materials, especially L 10
phased CoPt and FePt particles embedded in various non-magnetic matrices,
show considerable promise if the write-field limit can be alleviated.
New Magnetic Recording Media 219

1-1---1-1 -I Ru layer
1--
-I -\-
Traditional magnetic media AFC media

Figure 6.4 Schematic representation of traditional media and AFC media (arrow
denotes direction of moment) .

6. 3 Chemical Synthesis

6. 3. 1 Introduction
Advances in magnetic recording technology require new magnetic nanoparticle-
based media with uniformity in both particle size and particle magnetics
(Weller and Moser, 1999; Sun et al. , 2001). Solution phase chemistry-based
synthesis and self-assembly of monodisperse magnetic nanoparticles may offer
a suitable and convenient approach to such media. In these nanoparticle-based
films, grains (nanoparticles) are separated by a non-magnetic coating, which
minimizes exchange coupling between adjacent grains (Fig. 6. 5). With the
control of the assembly thickness, robustness, lateral dimension, and
magnetic easy axis orientation of the individual particle, such nanoparticle-
based media may present a magnetic recording medium possibly supporting
areal storage densities beyond Terabits per square inch.

Figure 6. 5 Schematic illustration of a self-assembled magnetic nanoparticle array. in


which each magnetic dot is surrounded by a non-magnetic medium.

One of the key challenges to realize this new paradigm is to make


monodisperse magnetic nanoparticles and nanoparticle arrays with control on
particle magnetics. Unlike the commonly used vacuum deposition techniques,
220 D.J.Sellmyer et a!.

which generally yield polydispersed magnetic grains, recent research progress


has shown that solution phase chemistry is an excellent approach to
monodisperse magnetic nanoparticle materials. It offers an important
homogenous nucleation step and facilitates isotropic growth of the nuclei
suspended in the solution. The nanoparticle dispersions can be stabilized
toward aggregation and oxidation with a layer of organic stabilizers. The
stabilized nanoparticle dispersion can be deposited on a solid substrate and
the solvent is allowed to evaporate. By controlling the concentration of the
particle dispersion and the solvent evaporation rate, 2-D or 3-D self-
assembled magnetic nanoparticle superlattices are formed. These well-
controlled magnetic nanoparticle arrays have shown spin-dependent tunneling
(Black et aI., 2000) and can support high-density magnetization reversal
transitions (Sun et aI., 1996). In this section, we focus on current chemical
synthetic approaches to monodisperse magnetic cobalt, iron and iron-platinum
nanoparticle materials.

6. 3. 2 Stabilization of Nanoparticle Dispersion

Magnetic nanoparticles experience strong van der Waals and magnetic


attractions. These attractions make the particle dispersion unstable and
particles prone to aggregation. As a result, it prevents one from using these
nanoparticles as building blocks to fabricate smooth and thin films required in
magnetic recording applications. To stabilize nanoparticle dispersion, a
stabilizing repulsive force must be present between two particles to counteract
the attractions and prevent particle aggregations. This stabilization can be
achieved by using either electrostatic repulsion or steric repulsion, as
illustrated in Fig. 6.6. Colloids remain stable with respect to aggregation only
if there exists a repulsive force of sufficient strength and range to counteract
the attractions. Coating the particles with ionic compounds will increase
electrostatic repulsion when particles are approaching each other. The well-
known gold sol is prepared by the reduction of [AuCI 4 ] - with sodium citrate
[(HOC(C02Na) (CH 2C02Na)2 , or CSHS07Na3]. Each gold particle is coated
with an electrical double layer of anions (such as Cs Hs O~- and CI-) and
cations (such as Na+ ). The electrostatic stabilization prevents gold colloids
from aggregation, leading to stable gold sol. Another way of protecting
particles from aggregation is to coat them with large molecules, such as
polymers or surfactants. The long chain hydrocarbons in polymers or
surfactants greatly increase the steric repulsion when particles get closer,
efficiently preventing them from aggregation.
New Magnetic Recording Media 221

(a)

(b)

Figure 6. 6 Schematic illustration of nanoparticle stabilization via coating of (a) IOniC


compounds or (b) long chain surfactants. The coating in (a) increases the electrostatic
repuision while the coating in (b) increases the steric repulsion among the particles.
These coatings can effectively stabilize the particle dispersion.

6.3.3 Chemical Synthesis of Monodisperse Nanoparticles


Organometallic-precursor decomposition and metal-salt reduction are two of
the most popular solution phase syntheses of magnetic nanoparticle
dispersions. Organometallic precursors, specifically metal carbonyls, are
generally used for the decomposition reaction, while a variety of metal salts
can be reduced to metal species by chemical reducing agents. Figure 6. 7
shows two examples commonly used to prepare cobalt nanoparticle materials.
It illustrates that decomposition or reduction of a cobalt precursor in the
presence of organic surfactants yields surfactant-coated cobalt nanoparticles
that are easily dispersed in a solvent to form nanoparticle dispersion.
6.3.3.1 Organometallic Precursor Decomposition
Cobalt nanoparticles have been prepared by thermally decomposing CO2 (CO)s
in refluxing toluene solution containing suitable polymeric materials, such as
methylmethacrylate-ethylacrylate-vinyl pyrrolidone terpolymer as stabilizers
(Thomas, 1966). The size of the particles was tuned from 2 to 30 nm by
varying reagent concentrations, temperatLire, and composition of the
polymers. Polymers containing a relatively large percentage of highly polar
groups, and higher polymer concentration, in general, resulted in smaller-
sized particles. It was also noticed that the copolymer of reasonably high
relative molecular mass (the order of 10 4 and greater) was unique in
222 D.J.Sellmyer et al.

'\1\1\I\,0

CoiCO)g I :>
Heat,-CO

(a)

(b)

Figure 6.7 Schematic illustration of cobalt nanoparticle synthesis via (a) cobalt carbonyl
decomposition and (b) cobalt salt reduction in the presence of organic surfactant
stabi Iizers.

stabilizing such colloids. Recently, decomposition of CO2 (CO)g has been


performed in the presence of lipid-type surfactants to make monodisperse
cobalt nanoparticles. Injection of CO2 (CO)g into hot solution containing oleic
acid / trialkylphosphine (R 3 P) (Murray et aI., 2001) and oleic acid /
trialkylphosphine oxide (R 3 PO) (Puntes et al. , 2001) has yielded variously
structured monodisperse cobalt nanoparticles. The decomposition of the cobalt
precursor occurs instantly upon injection, leading to the simultaneous formation
of many small metal clusters (nuclei). Continued heating allows steady growth
of these clusters into nanometer sized, single crystals of cobalt. The choice of
stabilizers allows average particle size to be controlled in the synthesis and a
series of monodisperse cobalt particles with sizes ranging from 3 to 12 nm have
been produced. Other cobalt organometallic precursors have also been used to
prepare cobalt nanoparticles. For example, under 3 bar dihydrogen,
decomposition of Co (Tj3-CgH13 ) (Tj4-CgH12 ) in the presence of
polyvinylpyrrolidone (PVP) at different temperatures 0, 20, 60 "C yields small
sized (trom less than 1 nm to about 1.5 nm) fcc cobalt nanoparticles (Osuna,
1996).
Decomposition of Fe(CO)5 is a common way to make Fe nanoparticles.
The synthesis has been performed in the presence of a variety of polymer
stabilizers (Griffiths et aI., 1979), sarkosyl-O (n-oleyoyl sarcosine)
(Wonterghem et al. , 1985; Wonterghem et al. , 1988; Bemtzon et al. , 1989;
Johansson et aI., 1993) and trioctylphosphine oxide/trioctyl phosphine
(TOPO/TOP) (Park et al. ,2000). Fe (CO)5 is also subject to ultrasonic
decomposition in solution phase to give Fe nanoparticles (Suslick et al. ,
1991 ; Cao et aI., 1995; Susl ick et al. , 1996). In contrast to the synthesis of
cobalt and iron particles, Ni (CO) 4 decomposition is rarely used to synthesize
New Magnetic Recording Media 223

Ni nanoparticles. This may be partly due to the high toxicity of Ni (CO)4


precursor that prevents its full use as a starting material. Recently,
decomposition of Ni(COD)z (COD = Cyclooctadiene) in methylene chloride in
the presence of PVP has been used to prepare Ni nanoparticles (Caro and
Bradley, 1997; Ely et al. ,1999).
6.3.3.2 Metal-Salt Reduction
In solution, metal ions can be reduced by a variety of reducing agents. The
reduction usually takes place instantly to give neutral metal species that further
grow into magnetic nanoparticles in the presence of stabilizers. This process is
illustrated in Fig. 6. 7b. A variety of inorganic reducing agents have been used
extensively to reduce metal salts to metallic particles. These include H--,
BH; -and BR 3 H -based reducing agents (Duteil et al., 1995; Bonnemann
et a!., 1990; Bonnemann et a!., 1992), and Li, Na, K naphthalides (Yiping
et a!. , 1990; Tsai and Dye 1993; Leslie-Pelecky et a!. , 1996; Leslie-Pelecky
et a!., 1998a; Leslie-Pelecky et a!., 1998b). Organoborohydride - type
reducing agents are found to reduce metal salts in solution phase to give high
purity magnetic nanoparticles with good monodispersity (Bonnemann et al. ,
1990; Bonnemann et al., 1992). We recently found that cobalt nanoparticle
materials can be easily prepared from cobalt chloride (CoCl z) reduction in the
presence of trialkylphosphine (R 3 P) and oleic acid (Sun and Murray, 1999;
Sun et a!. ,1999). Injection of an ether solution of superhydride (LiBEt3 H) into
a hot (200 "C, for example) CoCIz dioctylether solution leads to cobalt
nanoparticle materials. Reduction occurs instantly upon injection, leading to
the monodisperse cobalt nanoparticle materials. The choice of stabilizers and
reaction temperature allows average particle size to be controlled in the
synthesis and a series of monodisperse cobalt particle samples have been
produced. These organically stabilized cobalt particles are readily dispersed
in aliphatic, aromatic and chlorinated solvents, facilitating their assembly on a
sol id substrate.
Some organic compounds can be used as organic reducing agents to
reduce metal salts. The sonication of aqueous Cd+ and hydrazine yields
cobalt particle materials (Gibson and Putzer, 1995). A better-known organic
reduction process is called polyol process, in which organic diol is used as
both a reducing agent and a solvent (Fievet et a!. , 1989). In this process, the
metal salts are suspended in a liquid polyol (ethylene glycol, diethyleneglycol,
etc. ). The suspension is stirred and heated to a given temperature, which can
reach the boiling point of the polyol for more electropositive metals, e. g. ,
cobalt and nickel. The authors have recently extended the use of short
ethylene glycol to long chain diol, such as 1, 2-hexanedecanediol and 1, 2-
dodecanediol, in an almost stoichiometric manner in a high boiling ether
solvent. Stabilized hcp cobalt nanoparticle dispersion has been separated
(Murray et a!. , 2001).
Reduction of metal salts via electrochemistry is a clean way to prepare
224 D.J.Sellmyer et al.

magnetic nanoparticle materials. The synthesis employs a simple


electrochemical cell containing only a metal anode and a metal or glassy
carbon cathode. The electrolyte consists of organic solutions of tetra-
alkylammonium halide that also serves as a stabilizer for the produced
clusters. After application of a current, the anode undergoes oxidative
dissolution, forming a metal complex. The subsequent reduction of these
complexes leads to the formation of the desired particle dispersions (Reetz
and Helbig, 1994; Becker et aI., 1995). This approach allows the synthesis
of metal particles with narrow size distributions.

6. 3. 3. 3 Binary Metallic Magnetic Nanoparticles


The binary metallic magnetic nanoparticles, especially CoPt and FePt, are
promising candidates for future ultrahigh density magnetic recording because
they are magnetically hard and chemically stable. Compared to the synthesis
of elemental magnetic nanoparticles, the preparation of these bimetallic
magnetic nanoparticles is a more challenging task, as the techniques used to
make elemental metallic nanoparticles cannot be readily applied to synthesize
binary metal nanoparticles. The reduction or decomposition of two metal-
containing precursors often does not proceed in a way that facilitates the
formation of uniform binary metal particles. Instead, two kinds of metal
nanoparticle materials are often separated. As a result, chemical synthesis of
binary magnetic nanoparticles suitable for magnetic recording applications is
rarely reported. One way to form such binary nanoparticles is to mix two metal
precursors in a restricted environment, such as in reverse micelles. These
reverse micelles are formed by dispersing water droplets in the non-polar
solvent in the presence of surfactant molecules, such as sodium bis
(ethylhexyl) sulfosuccinate (AOT). They act as restricted nanoreactors.
Metal precursors are concentrated in each of the nanoreactors, facilitating
their growth into monodisperse nanoparticles. Two recent examples of using
these restricted environments to make CoPt and FePt nanoparticles include
sodium borohydride (NaBH 4 ) reduction of metal salts in cetyltrimethy
lammonium bromide (CTAB) based reverse micelles (Carpenter et aI., 1999;
Carpenter et al. , 2000) and ferritin-based protein shell (Warne et al. , 2001).
We recently reported that a combination of reduction of platinum
acetylacetonate [Pt(acac)2] and thermal decomposition of iron pentacarbonyl
[Fe (CO)5] at a high temperature (298C) could be used to prepare
monodisperse FePt nanoparticles (Sun et al. , 2000; Sun et al. , 2001). Oleic
acid and oleyl amine are proven to be a good ligand combination for FePt
particle stabilization. Oleic acid has long been used to stabilize varieties of
colloids including Fe nanoparticles (Suslick et al. ,1996). Alkyl amines, on the
other hand, are good stabilizing ligands for a noble metal such as Pt. The
structural similarity between oleic acid and oleyl amine provides a smooth
New Magnetic Recording Media 225

ligand shell around each FePt nanoparticle, facilitating superlattice formation.


The FePt nanoparticles are prepared by the combination of polyol reduction of
Pt(acac)2 and thermal decomposition of FeCCO)5 in the presence of oleic acid
and oleyl amine and can be easily dispersed into alkane solvent.
The composition of FePt nanoparticle materials is tuned by varying the
molar ratio of Fe CCO)5 and Pt Cacac)2. Based on 0.5 mmol each of the
PtCacac)2' oleic acid and oleyl amine, and 20 mL of dioctylether, the molar
amount of FeCCO)5 and the resulting Fe x Pt uoO - x ) are shown in Fig. 6.8 CSun
et al. , 2001). It can be seen that not all the FeCCO)5 contributes to the FePt
formation. FeCCO)5 has a low boiling point C103C). At reaction temperature
of 298 C , FeCCO)5 is actually in the vapor phase. The formation of this vapor
phase results in the slow decomposition of FeCCO)5 at a rate that matches the
reduction rate of PtCacac)2' The FePt nanoparticles are formed in a shorter
period of time. Therefore, the consumption of FeCCO)5 cannot be completed
on this synthetic time scale. As a result, 0.5 mmol of FeCCO)5 and 0.5 mmol
of PtCacac)2 yield Fe38 Pt62 , while 1.1 mmol of FeCCO)5 and 0.5 mmol of Pt
Cacac)2 lead to Fe56 Pt 44 nanoparticle materials. The FePt particle size can be
tuned from 3 to 10 nm by first growing 3 nm monodisperse seed particles in
situ and then adding more reagents to enlarge the existing seeds to the desired
size CSun et al. , 2000).

80

70

1
60
A;:
<l)

"'- 50
.S
"

0.5 1.0 1.5 2.0 2.5 3.0


Moles of Fe (CO\

Figure 6. 8 Compositional relation between Fe(CO)5 and x in FexPtnoO- x) based on


0.5 mmol of Pt(acac)2 (Sun et al. 2001).

6.3.4 Magnetic Nanoparticle Assembly

The monodisperse nanoparticles suspended in solution tend to form ordered


arrays after solvent evaporation. This assembly tendency is influenced by the
226 D. J. Sellmyer et al.

nature of the interactions exhibited among the stabilized particles. The


synthetic strategy to nanoparticle superlattices relies on a large number of
weak and non-directional interactions, such as ionic bonds, hydrogen bonds
and van der Waalss interactions to organize the particles-self-assembly. Such
self-assembled nanoparticles have been observed in a number of particle
systems (Murray et al. , 1995; Harfenist et al. , 1996; Harfenist et aI., 1997;
Pileni, 1997; Fink et aI., 1998; Korgel et al. , 1998; Vlasov et aI., 1999;
Jiang et al. , 1999; Wei et al. , 2000). The magnetic nanoparticles mentioned
above are very uniform in size, allowing them to self-organize readily into 2-D
and 3-D superlattices. Slow evaporation of a carrier solvent from particle
dispersion spread on a flat substrate allows well-organized superlattice
structures to be formed (Fig. 6. 9a). By varying the concentration of the
dispersion, both 2-D and 3-D particle superlattices can be formed. The TEM
image of a 2-D Co nanoparticle superlattice is shown in Fig. 6. 9b. The use of
concentrated dispersion of Co nanoparticles in higher boiling solvents like
dodecane allows slower evaporation at higher temperatures, producing well-
ordered multilayers (Fig. 6. 9c). Similar to cobalt particle assembly, when the
FePt nanoparticle dispersions are spread on a substrate, and the carrier
solvent is allowed to slowly evaporate, FePt nanoparticles self-assemble into
FePt particle superlattices. Fig. 6. 9d is a TEM image of a 3-D superlattice
assembly of 6 nm Feso Ptso particles.
The key challenge to the success of this self-assembly method is the
availability of large lateral dimensional arrays of well-controlled monodisperse
nanoparticles. A particular challenge for this self-assembly approach is the
preparation of robust thin particle assemblies for magnetic data storage
applications. Self-assembled magnetic nanoparticle arrays are usually held
together by very weak van der Waals and magnetic interactions. As a result,
the assembly is readily destroyed. For example, the assembly from the
particles protected by surfactants with long chain hydrocarbons can be re-
dissolved back into alkane, aromatic, or chlorinated solvents. To make a
mechanically robust particle assembly, thermal annealing or radiation
exposure is usually required to induce chemical reactions among the stabilizers
and interface reactions between the particles and the substrate. An alternate
approach to fabricate robust nanoparticle assemblies is to use template-
assisted assembly of nanoparticle materials (Sun et al. , 2001). This refers to
directionally assembling the particles on pre-treated substrates. These pre-
treated substrates are functionalized either with molecules, to facilitate the
chemical bonding between the particles and the substrate, or with porous
membranes, to facilitate organization of the particles into regularly arrayed
cores. Nanoparticles can be assembled in a controlled manner with robust
mechan ical properties.
New Magnetic Recording Media 227

n
(a) (b)

(c) (d)

Figure 6.9 (a) Schematic illustration of nanoparticle self-assembly via solvent


evaporation, and (b) TEM image of a 2-D assembly of 10 nm cobalt nanoparticles (bar =
28 nm), (c) TEM image of a 3-D assembly of 8 nm cobalt nanoparticles (bar = 80 nm),
and (d) TEM image of a 3-D assembly of 6 nm Feso Ptso nanoparticles (bar = 22 nm).

6. 3. 5 Conclusions
Many experimental results have proven that solution-phase chemistry is a
versatile approach to monodisperse magnetic nanoparticle materials. A
variety of chemical precursors are suitable for synthesizing such high-quality
magnetic nanomaterials via organometallic precursor decomposition and metal-
salt reduction. Numerous molecules, such as polymers and lipid-type
surfactants, can be used to coat nanoparticle surface to stabilize nanoparticle
dispersion. Stable magnetic nanoparticle dispersion is an excellent starting
material to build nanoparticle assemblies via particle's self-organization,
228 D. J. Sellmyer et al.

offering an easy approach to magnetic nanoparticle superlattices. Further


understanding and control of the rational assembly of these functional moieties
will be essential for any possible practical applications, as such control has not
yet been a reliable practice at the present time.

6.4 Self-Assembly

6. 4. 1 Introduction
Thermal stability is thought to be the ultimate fundamental limit to the magnetic
recording density, as conventional scaling leading to grain sizes smaller than
10nm (Weller and Moser, 1999). One way to overcome this limit is
patterning the media so that each grain represents one bit (White et al. ,
1997). E-beam and focused ion-beam lithography can reach features smaller
than 50 nm (Fischer and Chou, 1993; Wong et aI., 1999). However, these
processes are slow and cumbersome, so not suitable for mass production.
Optical lithography can be used to pattern a large area; however, it is well
known that the feature size is limited by the wavelength (Ross et aI., 1999).
Therefore, it is difficult to achieve periodicity smaller than 100 nm, which
corresponds to 65 Gb/in2 in areal density. Self-assembly has been recently
suggested to be a potential candidate for ultra-high density recording (Sun
et al. , 2000). Extremely small size (a few nanometer) and spacing (20-
30 nm), therefore high areal density ('" 1 teradot!in2 ) , can readily be
achieved by self-assembly, which is otherwise impossible by the conventional
lithography method. Self-assembled nanowire/dot arrays are usually highly
uniform in size and spacing, and quasi-ordered.
Techniques used to fabricate self-assembled magnetic nanowire/nanodot
arrays can be divided into two major categories: direct assembly of magnetic
dots (Sun et al. , 2000) and template methods. Basically, template methods
utilize the ordered structure of templates, and materials deposited
subsequently in these templates will reflect, more or less, the features of the
templates. Various self-assembled templates such as porous anodic alumina
(Zeng et al., 2000) and track-etched polycarbonate membrane (Fert and
Piraux, 1999) can be used. Recently, diblock copolymer (Albrecht et al. ,
2000) has been used as a template for self-assembly. Normally, the making of
nanowire arrays involves electrodeposition to fill the pores with magnetic
metals in templates because of the large aspect ratio, as opposed to
conventional vacuum technologies. This section will focus on the techniques of
porous anodic alumina template.
New Magnetic Recording Media 229

6. 4. 2 Porous Alumina Template Technique


The structure of templates: The top layer is porous alumina and the bottom is
aluminum; in-between is a nonporous barrier alumina layer. The pores are
parallel to each other and perpendicular to the film plane. Self-ordering
process occurs during the pore formation process. As a result, locally
hexagonally ordered pore arrays form. Two parameters: the center-to-center
spacing (Dc) and pore diameter (D p ) are used to characterize the
nanostructure of the template.
Sample preparation: The preparation of nanowire arrays in alumina
template involves the following major steps: electropolishing, DC anodization
and AC electrodeposition. The starting material, a piece of aluminum sheet is
electropolished at suitable conditions (e. g. , 60 V for 30 s, using 15% sulfuric
acid eletrolyte) to achieve a flat and smooth surface. The electropolished AI
foil is then anodized in acidic solutions under either constant voltage or
current. The commonly used acids for anodization are H2 S04' H2 C2 0 4 or
H3 P0 4 The combination of different anodization voltages (5 to 50 V) and
acidic solutions gives variable spacing and diameter of the templates. Multi-
step anodization/etching can be used to improve the regimentation of the
pores. Normally, AC electrodeposition is carried out for depositing magnetic
elements because of the dielectric nature of the barrier layer. For the
deposition of Fe, an electrolyte containing O. 1 mol/l FeS04 and O. 05 mol/l
H2 B03 is used; for the deposition of Ni and Co, FeS04 is replaced by NiS04
and CaS04' respectively. The deposition voltage is varied from 10 to 40 V,
depending on the anodization conditions. After electrodeposition, the opening
of the pores can be sealed by boiling in de-ionized water to prevent oxidation
of the nanowires. Alternatively, the AI layer and the barrier layer can be
chemically etched away from the template, a thin film of gold can be sputtered
at one side, and DC electrodeposition can be carried out. The advantage of
this method is the ease of controlling the deposition potential, especially when
multilayer or alloy nanowires are to be deposited. Instead of using AI foil, AI
film can be sputtered or e-beam deposited on n-type silicon (Crouse et al. ,
2000), and then the film is anodized. The advantage is that the resulting
template can be extremely flat.
The center-to-center spacing (Dc)' pore diameter (Dp) and ordering
depend on acids used and anodization voltages, and can be improved with
increasing anodization time. Typical values for Dc and Dpare 35 nm and 10 nm
at 10 V in 15% H2 S04 , 100 nm and 60 nm at 40 V in 3% H2 C2 0 4, and 200 nm
and 110 nm at 40 V in 4% H3 P0 4, respectively. Keeping Dc fixed, D p can be
further adjusted by immersing the as-anodized template in phosphoric acid for
different durations. The rate of pore widening depends on acids and
temperature, for example, when soaked in 1% H3 P0 4 at 25 t , D p is
230 D.J.Sellmyer et al.

approximately a linear function of time, and the widening rate is about


O. 5 nm/min. Macroscopic regimented area can be achieved by nano-
indentation before anodization (Masuda et ai., 1997). Figure 6.10 shows the
AFM image of a highly ordered alumina pore array.

0.75

~0.50

.-.
o
~

0.25
~R
. . . lB!l1o
0.50 0.75 1.00 ~m
0.25

Figure 6.10 AFM image of highly ordered alumina pore arrays.

Structural characterization of nanowires: Both the center-to-center


spacing and the diameter of the wires are controlled by the template. The wire
length is a function of deposition time, but also depends on numerous factors
including the template used (especially the thickness of the barrier layer), AC
electrodeposition voltage and frequency, temperature, concentration and pH
value of the solution.
As the wires are buried in the template, the signals from the X-ray are
weak. Transmission electron microscopy (TEM) has been the most powerful
means to characterize the nanostructure. Three methods for TEM specimen
preparation are available: plan-view, cross-section, and freed wires. Plan-
view and cross-section samples can be prepared by traditional dimpling and
ion milling techniques. Freed wires can be obtained by submerging the
specimen into a mixture of O. 2 mol chromatic acid and O. 4 mol phosphoric
acid. These wires are picked up by a metal grid coated with carbon film made
for TEM observation.
Figure 6. 11 shows representative TEM images and selected-area-
diffraction patterns of Fe, Co and Ni nanowires freed from the anodic alumina
template. Figure 6. 11 a, b are image and diffraction patterns of a bee-Fe
nanowire sample. The crystallite size is so small that it is not discernable in
the image, and the corresponding diffraction ring is very broad compared to
that of Co and Ni wires. However, at the opposite extreme, Fe nanowires
with single-crystallite size around 40 nm along the wire axis can also be
produced. For comparison, Ni nanowires show crystalline size around 10 nm,
New Magnetic Recording Media 231

with fcc structure, as seen in Fig. 6. 11 e - f. The nanostructure of Co wires is


more complicated. Crystallite size can be as large as a few 10 nm, and a
single wire consists of a chain of single crystallites; or the crystallites can be
extremely small, about 2 - 3 nm, and the cross-section of a wire consists of
5-10 grains. The Co nanowires consist of a mixture of hcp crystals and fcc
crystals. While fcc is a metastable phase for bulk Co, it is typically seen in Co
nanoparticles or ultra-thin films. Figure 6. 11 d shows the diffraction ring pattern
of fcc and hcp mixture of a typical Co nanowire sample. For samples that

(a) (b)

(c) (d)

(e) (f)

Figure 6. 11 Representative reflection images and TEM diffraction patterns of Ca) and
Cb) Fe. Cc) and Cd) Co and Ce) and en Ni nanowire.
232 D.J.Sellmyer et al.

contain mostly the hcp phase, the Co c-axis is randomly oriented, as revealed
by nanodiffraction patterns, and the crystalline size is extremely small ('" 2-
3 nm) (Zeng et aI., 2000). The size of the crystall ites of Fe, Co and Ni
nanowires, as well as the crystalline structure of the Co nanowires depends on
deposition conditions such as the AC frequency, pH value of the solutions, and
the chemical treatment of the as-anodized template before deposition.
Magnetic Properties of Nanowire Arrays: Typical hysteresis loops of Co
nanowire arrays are shown in Fig. 6. 12. The easy axis of the magnetization is
parallel to the wire and perpendicular to the plane. This originates from the
shape anisotropy of the wire. The local magnetocrystalline anisotropy of Co
grains are averaged out.

Hysteresis loops of Co nanowire arrays

1.0

0.5

:i 0.0
~
-0.5

-1.0
-10 -5 0 5 10
H (kOe)

Figure 6.12 Typical hysteresis loops of Co nanowire arrays.

The coercivity (He) of nanowires is strongly temperature dependent


(Wirth et al., 1998). This phenomenon can be attributed mainly to thermal
fluctuation effects, as well as a weak temperature dependence of intrinsic
properties such as saturation magnetization. He as a function of temperature
for Fe nanowires with different diameters is shown in Fig. 6.13. He decreases
rapidly with increasing temperature, being more prominent for thin wires.
Based on a phenomelogical model considering thermal activation over an
energy barrier (Zeng et aI., 2001), He (T) can be fitted by the following
equation:
(6.6)

where K a (n, Ha (n and Va are anisotropy, field and volume parameters,


whose physical meanings depend on the underlying model, and m varies
between 1. 5 and 2. o. It is found that only m = 1. 5 fits the data in the whole
temperature range. From the fitting, zero-temperature coercivity H co as a
function of diameter d w can be obtained, and is shown in Fig. 6. 14. For all
three materials, H co remains nearly a constant for small d w ; however, as d w
reaches a critical value, H co decreases with further increasing d w The
New Magnetic Recording Media 233

observed sharp transition shows some similarity with the transition from
coherent rotation to curling reversal. In perfect ellipsoids of revolution subject
to a field parallel to the long axis, magnetization reversal starts by coherent
rotation or curling, although there remains a remote possibility of the buckling
mode (Aharoni, 1996). The transition between the two modes depends on the

4.0
.: 5.5 nm
.: 10 nm
... : 12nm
... : 27 nm
+: 39 nm

2.0

50 100 150 200 250 300


T(K)

Figure 6.13 Coercivity as a function of temperature for Fe nanowires with varying diameters.

4.5
4.0 ... - r..... .: Fe
.: Co
3.5 \. "':Ni
,.,
0
3.0
2.5
..- ..", \.
\
.. '..
-'"
,, ' ....
~ 2.0
...-....
....
--
1.5
1.0 ...... ... ... ...
- .
....... _-----_!'
... -------
........
0.5
0 20 40 60

Figure 6. 14 Zero-temperature coercivity H dJ as a function of d w for Fe and Co. the


dashed lines are fits to Eqs. (6.8) and (6.9).

radius of the ellipsoid. The coercivity for curling reversal is

He = 2TIMs [ - ~; + (dw7do)2 ] (6.7)

where N z is the demagnetizing factor depending solely on the aspect ratio, k


is the shape factor, equal to 1.08 for an infinite cylinder and 1.39 for a sphere
(Aharoni, 1996). For infinite cylinders, coherent rotation occurs when the
diameter is smaller than a critical diameter de = k 1/2 do = 2. 08A 1/2 / M s ' do =
2A 1/2 / M s is the exchange length, and curling in thicker wires. de is calculated
to be 12, 15 and 27 nm for Fe, Co and Ni, respectively. Coercivity for d w
234 D. J. Sellmyer et al.

smaller than de should be equal to the shape anisotropy field 21TM s . Following
the d;/ dependence suggested by Eq. C6. 7), we have fitted our zero-
temperature coercivity data for Fe and Co to the expression:
H co = Ho ' d w < de; C6.8)
Hco = HI + CH o -H 1 )Cdc!dw)2, dw~de. C6.9)
The fits are shown by the dashed lines in Fig. 6.14, and the parameters are
HoCFe) = 4.1kOe, H,CFe) = 1.4kOe, deCFe) = 13.8nmandH oCCo) =
2.8kOe, H 1 = 0.8kOe, deCCo) = 14.5nm. AfitwasnotattemptedforNi
because de is in the range between 20 and 40 nm, which cannot be determined
due to insufficient data. The agreement between the fitted de values and
calculated values for Fe and Co is reasonably good. H co for thin wires are
O. 37,0. 33 and O. 32, respectively, that of the shape anisotropy field for an
infinite cylinder of Fe, Co and Ni.
Since coherent rotation and curling modes are delocalized CAharoni,
1996 ), the corresponding activation volume Can effective volume of
magnetization reversal) scales as the particle volume and diverge for long
wires. In reality, neither observed coercivities nor activation volumes support
delocalized reversal. The reason is that deviations from the limit of perfect
ellipsoids of revolution give rise to localized nucleation CSkomski et al. ,
2000). Due to the localized nucleation, the effective energy barriers are
reduced from shape anisotropy of an infinite cylinder, which in turn reduces
coercivity. The local ization also leads to effective volume of reversal
determined by the localization length, which is much smaller than the wire
physical volume. The fitting to Eq. C6.6) gives Vo parameter, which can be
viewed as an effective volume of reversal. In Fig. 6. 15, Vo as a function of d w
is plotted for Fe, Co and Ni. It can be seen that Vo is roughly proportional to
d~, and much smaller than the total wire volume.

60

50

a
~

40
.,
~<.)

S 30
C .: Fe
;:,.0
20 .:Co
.: Ni

10

o 5 10 15 20 25 30 35 40
d w (nm)
Figure 6.15 Effective volume of reversal Va as a function of d w for Fe, Co and Ni, and
the dashed lines are fitting curves assuming Va oc d~.
New Magnetic Recording Media 235

6.4.3 Self-Assembled Nanowire/Dot Arrays as Recording


Media-Merits and Challenges
It can be seen that self-assembly can achieve extremely small features of a few
nanometers. Self-assembled nanowires can have very uniform diameter,
center-to-center distance, and at the same time they can be well-ordered.
This is difficult to achieve by conventional thin-film technology and standard
lithography. Self-assembly only involves simple technology, which is cost
effective, and suitable for mass production. The magnetic properties of self-
assembled nanowire arrays show significant coercivity, high remanence ratio,
and narrow switching field distribution, which are all desired for high-density
magnetic recording. One of the most prominent features is probably that self-
assembled arrays can achieve areal density greater than 1 teradot/in2 .
However, as a technique still in development, self-assembled nanowire/
nanodot arrays are still far from application as recording media. One
of the major drawbacks is the electrodeposition technique involved.
Electrodeposition is suitable for depositing single element and simple alloys or
multilayers. For alloyed nanowires the controlling of the alloy composition and
deposition are complicated. For multilayer structures, electrodeposition is not
as versatile as vacuum deposition technologies such as sputtering. So far the
perpendicular anisotropy of nanowire arrays is mainly due to shape anisotropy;
no alloy with large magnetocrystalline anisotropy has ever been produced by
electrodeposition, which means that the coercivity is limited to only a few
kilooersted. A combination of the conventional thin film technology and self-
assembly may be a viable route to overcome these limits. For example,
substrates with self-assembled dots can be used to control the grain size of the
thin film media; nanopore arrays can be used as a mask, so that thin films can
be patterned into extremely dense arrays without time-consuming processes.

6.5 Concluding Remarks

Four approaches for new recording media development have been reviewed.
Each technique has its strengths and shortcomings. Sputtering is the most
mature and practical method for fabrication of commercial recording media at
the present time. However, the conventional grain size reduction has nearly
approached its limits. Thermal stability prevents further reduction of grain
size. Synthesis of new compounds with high anisotropy becomes critical for
this approach. Several compounds such as C05 Sm, Nd2Fe14B, FePt, and CoPt
have been obtained by sputtering followed by heat treatment. Heat treatment
236 D. J. Sellmyer et al.

can be replaced by laser surface heating which has the advantage of short
time, Iittle heating effect to the substrate, etc.
Chemical synthesis has shown great potential in its ability for obtaining
mono-distribution of grain size. Synthesizing textured media, establishing
repeatability, and synthesizing uniform media over a large area are challenges
for this method.
Self-assembled nanowires have the advantage of making use of shape
anisotropy. They can expand the superparamagnetic limit of spherical
particles. If used as a patterned medium, self-assembled nanowires represent
the highest achievable recording density. Breaking through in perfection of
ordering and making such media in a large area are the challenges.

References
Aharoni, A. Introduction to the Theory of Ferromagnetism. University Press,
Oxford (1996)
Albrecht, T. T., J. Schotter, G. A. Kastle, N. Emley, T. Shibauchi, L. K-
Elbaum, K. Guarini, C. T. Black, M. T. Tuominen and T. P. Russell.
Science 290: 2126 (2000)
Arnoldussen, T. C. , M. Mirzamaani, M. Doerner, K. Tang, X. Bian, J. Feng,
M. Gatherwright. Correlation of thermal stability and signal-to-noise ratio of
thin film recording media. Presented at TMRC, La Jolla, CA (1999)
Becker J. A. , R. Schafer, R. Festag, W. Ruland, J. H. Wendoff, J. Pebler,
S. A. Quaiser, W. Helbig and M. T. Reetz. J. Chem. Phys. 103: 2520
(1995)
Bemtzon, M. D., J. van Wonterghem, S. Morup, A. Tholen and C. J. W.
Koch. Phil. Mag. B 60: 169 (1989)
Bertram, H. N., H. Zhou and R. Gustafson. IEEE Trans. Mag-34: 1845
(1998)
Bertram, H. N. and J. G.Zhu. Solid State Physics Review 46 (1992)
Bian, B. , D. E. Laughlin, K. Sato and Y. Hirotsu. J. Appl. Phys. 87: 6962
(2000)
Bingham, D., M. J. Morgan and J. D. Cashion. Solid State Commun. 44:
517 (1982)
Birgeneau, R. J. , J. K. Kjems, G. Shirane and L. G. Van Uitert. Phys. Rev.
B 10: 2512 (1974)
Black, C. T. , C. B. Murray, R. L. Sandstrom and S. Sun. Science 290: 1131
(2000)
Bonnemann, H. , W. Brijoux and T. Joussen. Angew. Chem. Int. Ed. Engl.
29: 273 (1990)
Bonnemann, H. , R. Brinkmann, R. Koppler, P. Neiteler and J. Richter. Adv.
Mater. 4: 804 (1992)
Cao, X. , Y. Koltypin, G. Kataby, R. Prozorov and A. Gedanken. J. Mater.
Res. 10: 2952 (1995)
New Magnetic Recording Media 237

Cara, D. de and J. S. Bradley. Langmuir 13: 3067 (1997)


Carpenter, E. E., C. T.Seip and C. J.O'Connor. J. Appl. Phys. 85: 5184
(1999)
Carpenter, E. E., J. A. Sims, J. A. Wienmann, W. L. Zhou and C. J.
O'Connor. J. Appl. Phys. 87: 5615 (2000)
Crouse, D., Y.-H.Lo, A. E.Miller and M.Crouse. Appl. Phys. Lett. 76: 49
(2000)
Doerner, M. F., T. Yogi, D. S. Parker, S. Lambert, B. Hermsmeier, O. C.
Allegranza and T. Nguyen. IEEE Trans. Magn. Mag-29: 3667 (1993)
Duteil, A. , G. Schmid and W. Meyer-Zaika. J. Chem. Soc. Chem. Commun.
31 (1995)
Ely, T. 0., C. Amiens, B. Chaudret, E. Snoeck, M. Verelst, M. Respaud and
J.-M.Broto. Chem. Mater. 11: 526 (1999)
Fert, A. and L. Piraux J. Magn. Magn. Mater 200: 338 (1999)
Fievet, F. , J. P. Lagier and M. Figlarz. MRS Bulletin 29 (1989)
Fink, J., C. J. Kiely, D. Bethell and D. J. Schiffrin. Chem. Mater. 10: 922
(1998)
Fischer, P. B. and S. Y. Chou Appl. Phys. Lett. 62: 2989 (1993)
Fullerton E. E., D. T. Margulies, M. E. Schabes, M. Carey, B. Gurney, A.
Moser, M. Best, G. Zeltzer, K. Bubin, H. Rosen and M. Doerner. Appl.
PhyS. Lett. 77: 3806 (2000)
Gibson, C. P. and K. J. Putzer. Science 267: 1338 (1995)
Griffiths, C. H., M. P.O'Hora and T. W.Smith. J. Appl. Phys. 50: 7108
(1979)
Harfenist, S. A. , Z. L. Wang, M. M. Alvarez, I. Vezmar and R. L. Whetten.
J. Phys. Chem.l00: 13,904 (1996)
Harfenist, S. A., Z. L. Wang, R. L. Whetten, I. Vezmar and M. M. Alvarez.
Adv. Mater. 9: 817 (1997)
HoNo, K., Y.Maeda, J. L.LiandT.Sakurai. J. Magn. Magn. Mater. 110:
L254 (1992)
Inaba, N., Y. Uesaka and M. Futamoto. Compositional and temperature
dependence of Basic magnetic properties of CoCr-alloy thin films.
Presented at TMRC, La Jolla, CA (1999)
Ishikawa, A. and R.Sinclair. J. Magn. Magn. Mater. 152: 265 (1996)
Jiang, P. , J. F. Bertone, K. S. Hwang and V. L. Colvin. Chem. Mater. 11:
2132 (1999)
Johansson, C. , M. Hanson, P. V. Hendriksen and S. Morup. J. Magn Mag.
Mater. 122: 125 (1993)
Johnson, K. E. J. Appl. Phys. 87: 5365 (2000)
Kataoka, H., T. Kanbe, H. Kashiwase, E. Fujita, Y. Yahisa and K. Furusawa.
IEEE Trans. Magn. MAG-31: 2734 (1995)
Korgel, B. A. , S. Fullam, S. Connolly and D. Fitzmaurice. J. Phys. Chem. B
102: 8379 (1998)
Kubota, Y. , L. Folks and E. E. Marinero. J. Appl. Phys. 84: 6202 (1998)
238 D.J.Sellmyer et al.

Lee, L., D. Laughlin and D. Lambeth. IEEE Trans. Magn. MAG-30: 3951
(1994)
Lee, L. L., B. K. Cheong, D. E. Laughlin and D. N. Lambeth. Appl. Phys.
Lett. 67: 3638 (1995)
Leslie-Pelecky, D. L., M. Bonder, T. Martin, E. M. Kirkpatrick, Y. Liu, X.
a.Zhang, S.-H.Kim and R. D.Rieke. Chem. Mater. 10: 3732 (1998a)
Leslie-Pelecky, D. L. , X. a. Zhang, S. H. Kim, M. Bonder and R. D. Rieke.
Chem. Mater. 10: 164 (1998b)
Leslie-Pelecky, D. L., X. a. Zhang and R. D. Rieke. J. Appl. Phys. 79:
5312 (1996)
Liou, S. H., Y. Liu, S. S. Malhotra, M. Yu and D. J. Sellmyer. J. Appl.
Phys. 79: 5060 (1996)
Liou, S. H., S. Huang, E. Klimek, R. D. Kirby and Y. D. Yao. J. Appl.
Phys. 85: 4334 (1999)
Liu, Y., B. W. Robertson, Z. S. Shan, S. S. Malhotra, M. J. Yu, S. K.
Renukunta, S. H. Liou and D. J. Sellmyer. IEEE Trans Magn. MAG-30:
4035 (1994)
Lu, P. L. and S. H.Charap. IEEE Trans. Magn. MAG-30: 4230 (1994)
Luo, C. P., Z. S. Shan and D. J. Sellmyer. J. Appl. Phys. 79: 4899 (1996)
Luo, C. P. and D. J.Sellmyer. Appl. Phys. Lett. 75: 3162 (1999)
Luo, C. P., S. H.Liou, L.Gao, Y.LiuandD. J.Sellmyer. Appl. Phys. Lett.
77: 2225 (2000)
Madison, M. et al. J. Appl. Phys. 87: 4996 (2000)
Maeda, Y., S.Hirono, and M.Asahi Jpn. J. Appl. Phys. 24: L931 (1985)
Malhotra, S. S., Y. Liu, Z. S. Shan, S. H. Liou, D. C. Stafford and D. J.
Sellmyer. J. Mag. Mag. Mater. 161: 316 (1996)
Masuda, H., H. Yamada, M. Satoh, H. Asoh, M. Nakao and T. Tamamura.
Appl. Phys. Lett. 71: 2770 (1997)
McKinlay, S., Ph.D Dissertation, Stanford University (1999)
Mee, C. D. and Daniel E. D. Magnetic Storage Handbook. McGraw-Hili,
1996
Mirzamaani, M. and M. F. Doerner. IEEE Trans. Magn. MAG-32: 3638
(1996)
Murray, C. B., C. R. Kagan and M. G. Bawendi. Science 270: 1335 (1995)
Murray, C. B., S. Sun, W. Gaschler, H. Doyle, T. A. Betley and C. R.
Kagan. IBMJ. Res. & Dev. 45: 47 (2001)
Nakai, J., E. Kusumoto, M. Kuwabara, T. Miyamoto, M. R. Visokay, K.
Yoshikawa and K.ltayama. IEEE Trans. Magn. MAG-30: 3969 (1994)
Osuna, J. , D. de Caro, C. Amiens, B. Chaudret, E. Snoeck, M. Respaud, J.-
M. Broto and A. Fert. J. Phys. Chem. 100: 14,571 (1996)
Park, S. -J., S. Kim, S. Lee, Z. G. Khim, K. Char and T. Hyeon. J. Am.
Chem. Soc. 122: 8581 (2000)
Pileni, M. P. Langmuir 13: 3266 (1997)
Puntes, V. F., K. M. Krishnan and A. P. Alivisatos. Science 291: 2115
New Magnetic Recording Media 239

(2001)
Reetz, M. T. and W.Helbig. J. Am. Chem. Soc. 116: 7401 (1994)
Ross, C. A., H. I. Smith, T. Savas, M. Schattenburg, M. Farhoud, M.
Hwang, M. Walsh, M. C. Abraham and R. J. Ram, J. Vac. Sci. Technol.
B 17: 3168 (1999) and references therein
Shan, Z. S., S. S. Malhotra, S. H. Liou, Y. Liu, M. J. Yu and D. J.
Sellmyer. J. Mag. Mag. Mater. 161: 323 (1996)
Sharrock, M. P. J. Appl. Phys. 15: 6413 (1994)
Skomski, R. , H. Zeng, M. Zheng and D. J. Sellmyer. Phys. Rev. B 62: 3900
(2000)
Sun, S. and C. B.Murray. J. Appl. Phys. 85: 4325 (1999)
Sun, S. , C. B. Murray and H. Doyle. Mat. Res. Soc. Symp. Proc. 577: 385
(1999)
Sun, S. , C. B. Murray, D. Weller, L. Folks and A. Moser. Science 287: 1989
(2000)
Sun, S. , E. E. Fullerton, D. Weller and C. B. Murray. IEEE Trans. Magn. in
press (2001).
Sun, S. , D. Weller and C. B. Murray. In: Plumer M. , van Ek J. , Weller D.
eds. The Physics of High Density Magnetic Recording. Springer-Verlag,
Chapter 9 (in press)
Suslick, K. S., S.-B. Choe, A. A. Cichowlas and M. W. Grinstaff. Nature
353: 414 (1991)
Suslick, K. S., M. Fang and T. Hyeon. J. Am. Chem. Soc. 118: 11,960
(1996)
Thomas, J. R. J. Appl. Phys. 37: 2914 (1966)
Tsai, K. L. andJ. L.Dye. Chem. Mater. 5: 540 (1993)
Velu, E. M. T. and D. N.Lambeth. J. Appl. Phys. 69: 5175 (1991)
Vlasov, Y. A., N.Yao and D. J.Norris. Adv. Mater. 11: 165 (1999)
Warne, B. , O. I. Kasyutich, E. L. Mayes, J. A. L. Wiggins and K. K. W.
Wong. IEEE Trans. Magn. in press (2001)
Wei, Q.-H., D. M.CupidandX. L.Wu. Appl. Phys. Lett. 77: 1641 (2000)
Weller, D. and A. Moser. IEEE Trans. Mag. 35: 4423 (1999)
Weller, D. and M. F. Doerner. Annu. Rev. Mater. Sci. 30: 611 (2000)
White, R. L., R. M. H. New and R. F. W. Pease. IEEE Trans. Magn. 33:
990 (1997) and references therein
Wirth, S. , M. Field, D. D. Awschalom and r S. von Molnar. Phys. Rev. B 57 :
R14,028 (1998)
Wittig, J. E. , T. P. Nolan, R. Sinclair and J. Bentley. Mat. Res. Soc. Symp.
Proc. 517: 211 (1998)
Wong, J. , A. Scherer, M. Barbic and S. Schultz J. Vac. Sci. Technol. B 17:
3190 (1999) and references therein
Wonterghem, J. van, S. Morup, S. W. Charles, S. Wells and J. Villadsen.
Phys. Rev. Lett. 55: 410 (1985)
Wonterghem, J. van, S. Morup, S. W. Charles and S. Wells. J. Colloid
240 D. J. Sellmyer et al.

Interface Science 121: 558 (1988)


Xu, Y. F., J. P.WangandY.Su. J. Appl. Phys. 87: 6971 (2000)
Xu, Y. F. and J. P. Wang. IEEE Trans. Magn. MAG-36: (2000)
Yan, M. L., H. Zeng and D. J. Sellmyer(2001), (to be published)
Yiping, L. , G. C. Hadjipanayis, C. M. Sorensen and K. J. Klabunde J. Appl.
Phys. 67: 4502 (1990)
Yogi, T., T. A.Nguyen, S. E.Lambert, G. L.Gorman and G.Castillo. IEEE
Trans. Magn. MAG-26: 1578 (1990)
Yu, M. , Y. Liu, A. Moser, D. Weller and D. J. Sellmyer. Appl. Phys. Lett.
75: 3992 (1999)
Zhou, H. and H. N. Bertram. J. Appl. Phys. 85: 4982 (1999)
Zhou, H., H. N. Bertram, M. F. Doerner and M. Mirzamaani. IEEE Trans.
Mag-35: 2712 (1999)
Zeng, H. , M. Zheng, R. Skomski, Y. Liu, L. Menon and S. Bandyopadhyay.
J. Appl. Phys. 87: 4718 (2000)
Zeng, H., R. Skomski, L. Menon, Y. Liu, S. Bandyopadhyay and D. J.
Sellmyer. (2001) submitted to Phys. Rev. B

We thank C. P. Luo. S. H. Liou. R. Skomski. and S. Bandyopadhyay for collaboration and helpful
discussions. This work was supported by INSIC. NRI. CMRA. DOE. ARO. and NSF-MRSEC <DMR-
0213808).
7 Magneto-Optical Properties of Nanostructured
Media

Liangyao Chen, Songyou Wang and Roger D. Kirby

7. 1 Introduction

Magneto-optical phenomena have been known for more than 150 years. In
1845, Michael Faraday discovered the magneto-optical effect when he
observed that applying a magnetic field to a glass specimen caused the plane
of polarization of light transmitted through the glass to be rotated (Faraday,
1846). This rotation of the plane of polarization is called Faraday rotation. He
subsequently looked for similar rotation when light is reflected from a metal
surface in a magnetic field, but his results were inconclusive, evidently due to
scattering by surface imperfections. Thirty-two years later John Kerr
discovered this effect (Kerr, 1877, 1878) when examining the polarization of
light reflected from a polished electromagnet pole. The rotation in this case is
called Kerr rotation.
Three different orientations of the magnetic field with respect to the plane
of incidence of the light are commonly used to study the Kerr effect, and these
are shown in Fig. 7.1.

---
M

Polar Transverse Longitudinal


(a) (b) (c)

Figure 7. 1 Three high-symmetry configurations used in magneto-optic Kerr effect


measurements.

Figure 7. la shows the polar Kerr configuration in which the magnetization


is normal to the reflecting surface. The Kerr effect is largest in this case and it
is the only situation in which the Kerr effect occurs when the light is incident
242 Liangyao Chen et al.

normally on the surface. The polar Kerr effect is used in magneto-optical


recording (Reim and Schoenes, 1990).
Figure 7. 1b shows the longitudinal Kerr configuration in which the
magnetization M lies in the plane of incidence of the light. This is sometimes
referred to as the meridian Kerr effect. The longitudinal Kerr rotation is usually
smaller than that in the polar configuration by a factor of 3 or 4.
Figure 7. 1c shows the transverse Kerr configuration, where the
magnetization is perpendicular to the plane of incidence of the light. This is
sometimes referred to as the equatorial Kerr effect, and it leads only to a
change in intensity of the reflected light and not to a rotation of the plane of
polarization as the magnetization is changed. Any other situation can easily be
shown to be a combination of two or more of these configurations.
This chapter is devoted to demonstrating the importance of the polar Kerr
effect for studies of the magnetic materials, as well as to introduce the reader
to common methods of measuring magneto-optical Kerr rotation and ellipticity
as a function of Iight wavelength.

7. 2 Electromagnetic Theory of the Magneto-


Optical Effect

7. 2. 1 Faraday Rotation
We now consider light incident normally on the surface of magnetized sample,
with its wave vector propagating in the + z direction. We also assume that the
sample is magnetized in the + z direction. Phenomenologically, the inter-
action of light waves with a homogeneous material medium is described by the
dielectric tensor E. If the medium is isotropic, the three diagonal elements
will be degenerated, that is, E xx = E yy = E zz. When a magnetic field is
applied or the sample is magnetized in the z-direction normal to the surface,
the dielectric tensor will gain non-vanishing off-diagonal elements, E xy = - E yx ,
and the degeneracy of diagonal terms is lifted in the z-direction. Then the
dielectric constant tensor can be written as

~]
Exx E xy

E = - ~XY E/ . (7.1)
[
C E zz

The permeability tensor elements E Ij can be related to the conductivity tensor


elements CT Ij by the following expression:
. 411"
Er=
j
15 r-1-CTr
j W j
(7.2)
Magneto-Optical Properties of Nanostructured Media 243

where 5 ij = 1 for ; = j and 5 ij = 0 for the off-diagonal elements ; =1= j, w is


frequency of the light.
The conductivity tensor q is appropriate for most metallic magneto-optical
materials, where the dielectric constant diverges at low frequencies. All the
elements of ij and q ij are complex functions,

(7.3)

The dielectric response of the medium is described by Maxwell's equations

VXE=--.ldB }
c dt
(7.4)
VxH= ~(4TIJ+~~)
with V B = 0 and V D = 0 in the conducting medium.
From Maxwell's equations, it is easy to show that

(7.5)

where k is the wave vector in the z-direction, E is the electric field, and D is
the electric displacement field and is equal to E. Since k E = 0 and k = nw ,
c
we have
(ii 2 5 - E ) E = O. (7.6)

From Eqs. (7. 1) and (7. 6), in order to obtain a non-trivial solution, the
following secular equation should hold
ii 2 - 1(}( - xy 0
det xy ii 2 - 1(}( 0 =0 (7.7)
0 0 - zz

where ij are dielectric tensor elements.


Then the indices of the refraction for the two normal modes are given by
(7.8)

where the plus and minus signs correspond to eigenstates with right-circular
polarization (RCP) and left-circular polarization (LCP), respectively.
Equation (7. 8) shows that the off-diagonal dielectric element is essential to
magneto-optical activity, since RCP and LCP will be degenerated when xy
vanishes.
Substituting the expressions
(7.9)

into Eq. (7.8), and then setting I1n = n+ - n_ and 11k =


k + - k - , we
have the following relations between xy = lxy + ;2xy and I1n or 11k:
244 Liangyao Chen et al.

Elxy = kAn + nAk} (7.10)


E2xy = kAk - nAn
Linearly polarized light can be decomposed into left-and right-circularly
polarized components. As the linearly polarized light propagates into the
medium from z = 0 to z = L, the total phase difference between two modes
is given by

(7.11)

2lfk+ L
The intensity attenuation for these two normal modes are e--~- and
2lTk_ L
e--~-. Since the two circularly-polarized components are attenuated at
different rates, the transmitted light becomes elliptically polarized. The major
axis of the ellipse is rotated through the Faraday rotation angle eF relative to
the electric field axis of the incident light:
CA>L
e = 2"eCn+-
F n_). (7.12)

For most materials that are nearly transparent to show the Faraday effect,
the values of k are very small. Then the ellipticity that is defined as the ratio
of the minor to the major axis of the ellipse will be:
CA>L
EF =-2"CCk+-k-). (7.13)

Therefore, the Faraday rotation eF and ellipticity EF are proportional to the off-
diagonal elements of E xy :

eF - .IEF = -CA>L ~
2c -IE:
. (7.14)

The real and imaginary parts of off-diagonal elements of the dielectric


permeability tensor are related to the Faraday rotation and ellipticity as
follows:

(7.15)

7. 2. 2 Kerr Rotation
Now we turn our attention to the magneto-optical effect for reflected light. We
restrict our calculations to the polar Kerr effect, where the incident light is
normal to the sample surface. The optical reflectance coefficient r is
determined by the Fresnel equations. At normal incidence, the reflected
Magneto-Optical Properties of Nanostructured Media 245

amplitudes r + and r - for right and left circular polarizations are

r = I r Ie i8 = n + ~ k - 1 (7.16)
n+ Ik+ 1
and the reflected light intensities R for the two polarizations are

R - 2 _ (n- 1)2 + k~
(7.17)
- r- (n+ 1)2 + k~
The complex polar Kerr angle k is defined as:
k = ek - iEk (7. 18)
In the polar geometry, the magnetization is perpendicular to the surface and
along the direction of light propagation, i. e., in the + z direction. The
eigenmodes are right-and left-circular polarizations, and the components of k
are expressed in terms of the Fresnel coefficients (Eq. (7. 16)) as

(7.19)

(7.20)

where e is phase angle of reflection coefficient related to the right (+) and
left (-) circular polarized light, respectively.
It should be noted that after reflection, the direction of light propagation is
reversed, traveling now along the - z direction. This implies that for right-
circularly polarized light ( + ) the electric field vector will rotate
counterclockwise when viewed along the light propagation direction. The sign
of the Kerr rotation angle has been chosen, and the ellipticity Ek will be
positive at the photon energies where the right-circular polarization field is
more strongly attenuated than the left-circular polarization field. This is easily
seen by writing the phase and ampl itude of r (Eq. (7. 16)) .

2k ) (7.21)
e = arctan ( n~+ k~- 1

- J(n~+ k~- 1)2 + (2k )2


I r I - (n+ 1)2 + k~ (7.22)

Using Eqs.(7.21) and (7.22) and making the approximations:


sin (e + - e_) =( e+ - e_), cos ( e+ - e_) 1, and (r + = - r _ ) 2 2 r + r _ ,
the following expression can be derived,

(7.23)

These approximations should lead to small errors in most instances, since in


most situations, the Kerr rotation will not be larger than few degrees. SLbstituting
the Fresnel equations, Eq. (7.16) into Eq. (7.23), we find the following
246 Liangyao Chen et al.

relationship between the complex Kerr angle and the complex indices of
refraction:
n+- n_
4>k = Bk - iEk = i -:-_-_,.--- (7.24)
n+ n_- 1
Using the relation between the dielectric constant E and nand k ,

E = n~ = (n+ ik)2 = 1 - i 41Ta (7.25)


w
where a can be related to a xx and a xy ,

a= a xx ia xy . (7.26)
Therefore, we have the relation between it and conductivity tensor a from
Eqs. (7.25) and (7.26)

(7.27)

so that we obtain the complex Kerr angle as a function of Exx and Exy ,

- -
n+-n_ xy-a E
4>k = Bk - iEk = i -=_-----:_=__- = - - = _ _--=....:xLy_ _ (7.28)
n+ n_- 1 axxn (1 - Exx ) ..IE:
-
Here, we have used the approximation n ="21(n- +
-
+ n -)
- - -2
and n + n _=n .
Eq. (7.28) indicates that magneto-optical Kerr effect will be large when the
denominator is small. For a metal, this occurs at about the plasma frequency
where the real part of E xx can be near zero or close to 1, even with a moderate
value of E xy .
From Eq. (7.28) and E xx = (n + ik)2, the real and imaginary parts of
E xy (or a Xy) can be expressed in terms of a linear combination of the Kerr
rotation and ellipticity, as in the following relations:

(7.29)

or

alxy = 4~/BBk - AEk)}


(7.30)
a2xy = 4~ (AB k + BEk)
where A and B are expressed by
A = n 3 - 3nk 2 - n = neE! - 1) - kE2 }
(7.31)
B =- k 3 + 3n 2k - k = keEl -1) + nE2 .
Magneto-Optical Properties of Nanostructured Media 247

Thus from the experimentally measured values of the Kerr rotation and
ellipticity, one can calculate the real and imaginary parts of the off-diagonal
elements of permeability tensor (or conductivity tensor), provided that optical
constants nand k are known.
To fully characterize the magneto-optic properties, it is essential to
measure the spectra of E xy (or a Xy) with high precision. As is shown in the next
section, E xy is directly connected to the properties of the electronic band
structure of the material.

7. 3 Microscopic Models

Magneto-optical phenomena are due to the interaction of light with the matter
and therefore, in principle their study requires quantum mechanics. Hulme
( 1932) was first to propose a tentative quantum explanation of magneto-
optical effects, and his analysis was then extended by Kittel (1951) and
Argyres (1955). They pointed out that, in general, magneto-optical effects
will be proportional to

~ (S, x \TV) P, (7.32)


I

which is the spin-orbit energy of an electron with spin S and momentum p


moving though the electric field - \TV inside the medium, summed over all
electrons in the medium.
Assuming this proportionality to spin-polarization and spin-orbit coupling,
Bennet and Stern (1965) and Erskine and Stern (1973) developed expressions
for a xy which allow quantitative calculations, taking into account both
interband and intraband transitions of magnetic electrons in the medium.

7. 3. 1 Interband Transitions
The off-diagonal element of the optical conductivity a xy evaluated by quantum
mechanics is given by

a xy = (2we~2)~~[I(lT-)"" 1
2
-I(lT+)"., 12J/[W~n-(w+in2J
(0) (u)

(7.33)
where the operators = IT x ilT yare linear combinations of the kinetic
IT

momentum operator defined by IT = P, + (_11_) s,x \TV ( r I) which includes


I
4mc 2
spin-orbit interactions, and nand m are summation indices over the occupied
248 Liangyao Chen et al.

(n) and unoccupied (m) states.


Measuring the magneto-optical spectrum over a wide range of frequencies
lets one determine 0'2xy' the imaginary part of 0' xy. Then O'''y can be
calculated from the Kramers-Kronig relation. By taking a limit as 1 - 00,
Eq. (7.33) becomes

0'2xy = (4:~~2)~ ~[I(-rr-)mn 1


2-I(lT+)mn 1
2
]<5(w-w mn ).
n m
(0) (u)

(7.34 )

The spectrum of 0' xy can be obtained from first-principles band-structure


calculations by evaluating the transition matrix elements of right (+) and left
( -) circular polarization (IT ) mn for all k -points in the Brillouin zone, taking
into account the spin-orbit interaction. Such calculations have been performed
for Fe (Singh et aI., 1975); Ni (Wang and Callaway, 1974), Pd (Yaresko
et al., 1998) and Pt (Uba et al., 2000), and they show fairly good
agreement with experimental results. Recently, better relativistic calculations
have been given by Ebert's group for the 0'2xy spectra of PtFe and PdFe alloys
(Ebert and Akai, 1990).
However, many of the band structure calculations only take into account
the density of states, or at best the joint density of states, and do not include
transition matrix element effects. In this case, Eq. (7. 35) provides a
convenient way to evaluate 0'2xy:
2
0'2xy = 4wm
lTe
2 11
{F-11m J-11m - F+11mJ+11m } (7 35)

where F;m is the averaged transition matrix for elements of right (+) circular
polarization RCP( +) and LCP (-) circular polarization LCP( -), and J;", is
the joint density of states. Their product is given by

F:::mJ;" = 8:3 f(
I (11 t IlT 1 n t) 1
2
+ I (m t IlT I n t) 1
2
) <5 (w - W 11m) d3 k
(7.36)

where t and ... indicate the polarization of spin up and down, respectively.
The balance between the contribution of RCP and LCP in Eq. (7. 35) can be
approximated by energy-derivatives of joint density of states, which provides
a satisfactory explanation of MO spectra in Cr 3 Te4 (Sato et aI., 1992) and
PtMnSb (Takanashi et aI., 1991).

7. 3. 2 Intraband Transitions
It can be shown that the intraband transitions may contribute to 0'2xy with two
different frequency dependencies. The theory gives a proportionality of 0' xy to
the conduction-electron concentration N via the plasma frequency
Magneto-Optical Properties of Nanostructured Media 249

4-rrNe 2
w~ - -m-'- (7.37)

where e is the electron charge and the m' is the effective mass of the
electron, and to the spin polarization

(1 cond =
nt - n~ (7.38)
nt + n~
where n t and n ~ are the numbers of conduction electrons with the spin
moment parallel and anti-parallel to the total moment. In addition, a second
term is proportional to the strength of spin-orbit coupling via the ratio of
maximum macroscopic dipole m()ment Po and the product of electron charge
and the Fermi velocity evo.

w~(1
(W ) = 4-rr {_ Q + Po [1- iw(y+iw) ]}
(1xy cond Q2+(y+iw)2 evo Q2+(y+iw)2
(7.39)
where y = 1/ T and Q are the scattering and skew-scattering frequency,
respectively. T is relaxation time. The absorption part of the free-electron
contribution is then

(7.40)

It is important to note that the two terms in Eq. (7.40) have opposite signs.
The first term in Eq. (7.40), with its (1 2xy OC W -3 frequency dependence
in the high-frequency limit, gives the same frequency dependence as the
classical Drude treatment of the free electrons. This term dominates if the
damping is small enough to allow a coupled plasma frequency W p well below
the lowest energy interband transition. As we will see in the next section, it
may be viewed as a splitting of the plasma edge for the two circular
polarizations and leads to very large magneto-optical effects if the plasma
edge in the optical reflectivity is steep.
In the classical Drude treatment one starts with the equation of the
electron motion under electric and magnetic fields:
2
d r + . dr + dr H E (7.41)
m df m y dt ello dt x =- e .

Assuming an e''''t time dependence, the dielectric constant for RCP and LCP is
found to be:
w2
E = 1+ p
w(-wwe+iy)
(7.42)

where We is the cyclotron frequency


250 Liangyao Chen et al.

epo H
W
C
=m*'
-- (7.43)

Using Eqs. (7.2), (7.25) and (7.26), we obtain


2
-IWeW p
(7.44)
w[(iw + y)2 + w~J
and

(7.45)

Equation (7. 45) has a form that is equivalent to the first term of Eq. (7. 39)
(except for the conduction-electron spin-polarization O"coocj shown in
Eq. (7.39)) if We is identified with the skew-scattering frequency Q.
As will be seen in a later section, the experimental Kerr effect
measurements on paramagnetic silver (Deng et al. , 2001) and Co-Ag (Wang
et al. , 2000a) granular films show the characteristics expected from plasma-
edge splitting.

7. 4 Magneto-Optical Measurements

Many different experimental techniques have been used to measure magneto-


optical spectra (Suits, 1971; Sato, 1981; Katayama and Hasegawa, 1981;
Chen et al. , 1997; Manens et al. , 1984; Mansuripur, et al. , 1990). The most
common techniques employ: CD a vibrating polarizer; (2) a rotating polarizer;
@a Faraday modulator or @a modulation of the phase retardation. Since
techniques CD - (2) utilize modulations of linear polarization, they are not easily
used to measure the ellipticity spectra. Mansuripur et al. (1990) presented a
method to measure both the Kerr rotation and ellipticity, but there is a sign
uncertainty in determining the rotation angle. We have used a wavelength-
scanning Fourier transform magneto-optical method to measure Faraday and
polar Kerr rotation ellipticity. This method employs a uniformly rotating
analyzer combined with an achromatic quarter-wavelength retarder. Recording
the time-dependent output signal and subsequently taking a Fourier transform
to obtain both the amplitude and phase allows both the rotation and the
ellipticity to be determined (Chen et al. , 1997).

7. 4. 1 Experimental Principles
We assume an isotropic material with magnetization M in the + z direction
(normal to the sample surface). The incident light also propagates in the + z
Magneto-Optical Properties of Nanostructured Media 251

direction and is linearly polarized along the x direction. Then, the electric
field can be decomposed into two right- and left-circularly polarized
components, E, and E i. e. ,
"
(7.46)
If the material shows a Kerr effect, the field E' that is reflected by the sample
will be elliptically polarized, i. e. ,
(7.47)

where r + and r _ (r = I r I eI8 ) are the reflection coefficients for the right-
and left-circularly polarized light, respectively. The small y component of the
electric field E~ arises purely from magneto-optical effects. Therefore, the
ratio of E'y to E: is
E'y _ i( r + - r _)
(7.48)
E~ - f++ r_

Since the Kerr rotation and ell ipticity, fJ k and Ek' have the forms

(7.49)

the complex Kerr angle 4>k becomes

(7.50)

where E xx and E xy are the diagonal and off-diagonal components of the


dielectric tensor, respectively.
Therefore, there are two ways to enhance the Kerr effect. One is to find
a material intrinsically having a higher E xy value, resulting in a larger
component E~. The other is to reduce the value of E:, which is proportional
to the pseudo reflectance coefficient of the material. As can be seen from
Eq. (7.50), if the pseudodielectric constant E xx close to 0 or close to 1, the
Kerr effect will be significantly enhanced. However, the reflected light
intensity will be seriously reduced, because the reflectivity will close to zero
under this condition.
After the light is reflected from the sample, a polarization analyzer is used
to detect the polarization state of the light. The azimuthal angle of the analyzer
is A with respect to the x-axis. The electric field E emerging from the
analyzer is
E = ExcosA + EysinA (7.50
so that the light intensity I can be written as

I::::::: ~2[f~+ f~+ 2f+ f_ cos(2A - 2fJ k ) ] . (7.52)

Although, in principle, both fJ k and Ek can be determined by the analysis of the


252 Liangyao Chen et al.

Fourier transform, there is a technical difficulty in determining the ellipticity


Ek' especially its sign, since the background signal is included in the dc
component along with the squared terms in the equations. The method can be
improved by placing an achromatic compensator in the optical path between
the sample and analyzer. Letting a be the retarding phase angle of the
compensator, Eq. C7. 47) is the modified to be
C7.53)
The light intensity is then easily calculated to be
I =1
0 + 11 cosC2A) + I z sinC2A) C7.54)
where

10 = f Cr~ + r~) }
11 = fC2r+ r_ cosC20 k C7.55)
I z = f[ - Cr~- r~)sina + 2r+ r_ sinC20k)cosa]
Here, f is a constant depending on the characteristics of the photon detector.
Since r+~ r_~ r and Ok is small, the ratio of I z to 11 is
Iz -Cr+-r_)sina+rsinC20k)cosa 2CO . )
-I ~ C20 ) ~ kcosa - EkSlna .
1 rcos k

C7.56)
Thus, Ok and Ek will equal I z /1 1 for a = 0 Cthe compensator is removed from
the optical path) and for a = 90' (the compensator is placed in the path),
respectively.
The three light components in Eq. (7. 54) can be calculated from the
Fourier transform:

C7.57)

If the compensator has a phase error due to optical imperfection, i. e. ,


a = 90' Aa C7.58)
there will be an error in determining Ek'

Ek -- 21; - ekL>a.
lz + A C7.59)

Assuming that the magnitudes of Ok and Ek are of the same order and Aa
is less than about 1', the relative error in Ek will be less than about 1% .
The method also can be used to measure the Faraday effect. In the
Magneto-Optical Properties of Nanostructured Media 253

standard Faraday configuration, the transmitted field has a similar form to


Eq. (7.47)
(7.60)

where t + and t - are the transmittance coefficients for the right and left
circularly polarized light, respectively.
Unlike the Kerr effect, the Faraday parameters depend on both the
magnetic field and the sample thickness. For very thin film samples or small
Faraday effect, Eq. (7.56) is still valid for measuring the Faraday effect:

~: ::::::::; 2(8FcosO' - EFsinO'). (7.61)

If the sample is quite thick, resulting in a large Faraday rotation, then 8F and
EF are determined by assuming that t + ::::::::; t - ::::::::; t:

/2 sinO'
1;::::::::; tan(28F)cosO' - 2EF cos(28 F)' (7.62)

Here 8F can be measured precisely at 0'= O. Error in measuring EF' however,


will occur since AO' and 8F are correlated. An appropriate error reduction
procedure will be needed in the data analysis.

7. 4. 2 Description of the Experimental System


The Kerr and Faraday optical systems, mounted on an optical table, are
shown schematically in Fig. 7.2. The sample is placed between the poles of a
13 kOe electromagnet, and transmitted light is measured with the Faraday
analyzer (right-hand side of the apparatus), while reflected light is measured
using the Kerr analyzer (left-hand side of the apparatus). A 150-W Xe short-
arc lamp provides a continuum light source to cover the 1.5 - 4.5 eV (830-
300 nm) spectral range. Quasi-monochromatic light is provided by a high-
throughput quartz prism monochromator. The light is conducted to the Faraday
and Kerr measurement systems via a pair of fused-silica fiber optic cables, as
shown in Fig. 7.2. This arrangement makes the system alignment easier. The
polarizer and analyzers are improved Glan-Foucault calcite prisms, especially
0
designed to provide beam deviation angles of less than O. 01 The prisms

have an air gap between the prisms, so that the spectral response can be
extended to the ultraviolet range. A front-surface mirror is used to pass light to
the sample through a small hole in the pole face of the magnet. Both the mirror
surface and the azimuthal direction of the polarizer are carefully adjusted to
make them perpendicular to the plane of incidence of the light. Therefore,
highly pure polarized light will be incident on the sample. The angle of
0
incidence is less than 2 and fixed. The magnetic field can be swept over a
range of 13 kOe under computer control.
254 Liangyao Chen et al.

Figure 7. 2 Schematic diagram of the experimental system for the spectral measurement
of Kerr and Faraday effects. 1: monochromator; 2: reflector; 3: beam apertures;
4: polarizer; 5 and; 6: adjustable front surface mirrors; 7: sample; 8: quarter-
wavelength retarder; 9: stepping motors with hollow shafts; 10: polarization analyzers
mounted directly on motor shafts; 11: light-shielding box; 12: vibration-isolated optical
table.

To avoid mechanical transmission problems, the analyzer is directly


mounted onto a hollow shaft of a high-resolution stepping motor (1000 steps
per revolution). The achromatic quarter-wavelength retarder (item 8) is a
fused-quartz rhomb-type-F prism, which has superior quality, with a retarding
error of less than 0.3 in the 200 - 800 nm wavelength range, and with very
small center transmitted beam displacement. The retarder can be moved into
or out of the optical path by a stepping motor, depending on whether Kerr
rotation or ellipticity is being measured. The Fourier transforms are then
performed using Eq. (7.57) to obtain either the polar Kerr or the Faraday
parameters through Eq. (7.56) and Eq. (7. 61>, respectively.
Since the Kerr effect is an odd function of the magnetization. The complex
Kerr angle cPk is measured by taking the average of two measurements in
positive and negative magnetic fields, respectively, as

(7.63)

This procedure cancels system errors that are not magnetically dependent,
such as the errors from the uncertainty in determining precisely the initial
azimuthal angle of the analyzer, the phase delay due to the electrical circuits,
any slight optical misalignment of the system, the slight differences of the
Magneto-Optical Properties of Nanostructured Media 255

sample position in each measurement, etc. The same procedure is used in


measurements of the Faraday effect.
Since Eqs. (7.56), C7. 57) and C7. 61) are quite straightforward, the
system can easily be used to measure both the absolute Kerr and Faraday
values over a wide range from one-hundredth to a few hundred degrees (Chen
et aI., 1996).

7. 5 Magneto-Optical Spectra of Magnetic and


Non-Magnetic Metals

7.5.1 The Magneto-Optical Effect in the Noble Metal Ag


Magneto-optical Kerr effect spectroscopy is a useful tool for the study of
magnetic properties and electronic structure of magnetic materials. Some
specific examples are now considered for the verification of the macroscopic
formulas presented in Section 7. 3 .
0.02

--.
'-"
of -No.1
- - - No.2
........ Ref.
-0.04

3.6 3.8 4.0 4.2


E(eV)
(a)

0.04

--. 0.02
'-"
~
0

E(eV)
(b)
Figure 7. 3 Experimental spectra of Kerr effect of Ag thin films, - - prepared by
thermal evaporation,----- prepared by magnetron sputtering(taken from Phys. Rev. 183,
(1969)664).
256 Liangyao Chen et al.

Figure 7. 3 displays the experimental Kerr effect of paramagnetic silver


for applied magnetic field of 10 kOe, which is a typical example to show the
large Kerr effect that can occur due to plasma-edge splitting (Deng et al. ,
2001). Figure 7.4 gives the detailed photon energy dependence of ek and ek
The experimental results can be successfully modeled using Eqs. (7.28) and
(7.44). The origin of the Kerr effect in a paramagnetic sample is that the
spins of the d-electrons of Ag are partially aligned by the external magnetic
field, and the right- and left-circular polarizations have different plasma
frequencies. On the other hand, both A / (A 2 + B 2 ) and B / (A 2 + B 2 ) as
determined in Eq. (7.29) are small, and B changes its sign near the plasma-
edge of Ag (Fig. 7.5). Obviously, large Kerr effects are generated in the
vicinity of the plasma resonance. Similarly, large Kerr peaks in TmS and
TmSe have also been attributed to an exchange splitting of the plasma edge
(Reim et aI., 1984). This exchange splitting will result in a resonance-like
pec:i1< of the Kerr effect at the plasma frequency (Feil and Haas, 1987) .

- W c =O.lmeV
~
- - W c =0.12meV
'-' -0.02 ....... w c =O.13 meV
'""" _.- wc =0.135 meV H
(;
_ .. - w c =0.14 meV

-0.04
2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

0.02 ~

'-'
0
c:e
-0.02

2.5 3.0 3.5 4.0 4.5


E(eV)
(b)

Figure 7.4 Calculated e


k and Ek spectra for Ag.

7.5.2 Magnetic Granular Films


Magnetic granular solids have received considerable attention in the past few
years (Xiao and Wang, 1994). In these composite materials, ultrafine
Magneto-Optical Properties of Nanostructured Media 257

01----------=""'"'-
-0.5
--No.1
---No.2
-1.0
......... Ref.
:f
~
-15
.
"'<: -2.0

-2.5 L -_ _----' --'- -L._ _----'


2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

2.0
1.5
1.0

~ 0.5
,t
0
~ -0.5
-1.0
-1.5
2.5 3.0 3.5 4.0 4.5
E(eV)
(b)

Figure 7. 5 A / ( A2 + B2 ) and B / ( A 2 + B2 ) spectra calculated from the measured


dielectric constants of Ag.

magnetic particles of a few nanometers in size are embedded in a metall ic or


insulating matrix by certain synthesis processes. Because of their
nanostructure and the capability to control material and geometrical
parameters, granular materials possess different and sometimes enhanced
magneto-optical properties as compared to their bulk counterparts. In
particular, a giant magnetic coercivity (He) has been achieved in Fe-Si02
granular films, with He enhanced by as much as three orders of magnitude over
the bulk value (Xiao and Chien, 1987). It has also been discovered that
metallic granular alloys, such as Co-Cu, Co-Ag, and Fe-Ag, exhibit giant
magnetoresistance effects (Berkowitz et a!., 1992; Xiao et a!. , 1992, 1993;
Xiong et al., 1992; Wang et al., 1993). These developments in granular
solids call for more systematic studies of their magnetic properties. In recent
years, the magneto-optical Kerr effects in Ni-Si02 , Co-Cu, Co-Ag and Fe-Ag
granular films have been studied (Zhang et al. , 2001; Wang et al., 1999,
2000a, 2000b). The size of the magnetic particles and their magnetic
interactions can be at least partially controlled through variations in film
composition and post-deposition thermal treatment. This will clearly affect the
258 Liangyao Chen et al.

magneto-optical properties of the materials.


We now report the magneto-optical properties of several composite films.
All of the films were fabricated by sputtering at room temperature from a
composite target. Some samples were annealed under vacuum for one hour at
different temperatures in order to vary the nanostructure. The complex
dielectric function E = El + iE2 was measured at room temperature using a
scanning ellipsometer (Chen et al., 1994). The complex polar magneto-
optical Kerr spectra were obtained using an applied field of 10 kOe. All
measurements were made ex situ at room temperature.
7. 5. 2. 1 Ni-SiOz Granular Films
Figure 7. 6 shows the spectra of polar magneto-optical Kerr rotation and
ellipticity in Nix (Si02) l-x (Zhang et al. , 2001). It can clearly be seen that the
magnitudes of the Kerr spectra increase with increasing Ni content. A broad
peak in 8k is located at 2.5 eV for the x = O. 25 film, and it shifts to 3. 2 eV
for the x = O. 84 film. Similarly, the broad peak in Ek shifts from 3. 5 to

0.20

0.15
_-_? .
.-.
.' ~.,;;::::~.~-;:-....:'"
0.10

0.05
.. x
'..
"',
"' I
, 1'..
::(.
\\

......' " \
I: 0.52 ',~ . " \
o 2: 0.58 , ....."'\
-0.05 ~: ~:~~ "'''-_.~.~~
5: 0.84
-0.10 6: 0.65 (annealed)
-0.15 L - _ - - L - _ - - - - ' -_ _L.-_--'---_--L_---'
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

0.20

0.15

0.10

0.05
~
...
to
0
-0.05

-0.10
L-_--'----_--L_----J'--_-'-_-'--_----J
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(b)

Figure 7. 6 Measured Ok and Ek spectra of Nix (Si02 )o-x) granular films with
different Ni content.
Magneto-Optical Properties of Nanostructured Media 259

4.2 eV for the same films. Figure 7. 7 shows the complex dielectric function
(1 and 2) spectra for the same Nix (Si02 )l-x films. The real part of
dielectric function, l ' decreases with increasing Ni content. By x = 0.84,
the sign of , becomes negative in the low-energy region, and changes from
positive to negative at 3.6 eV. However, the imaginary part 2 increases with
increasing Ni content and becomes more like the spectra of pure Ni.

4
.... '-
3
,, >.. . . .
..... '-
-----------~
_ 1 ""- "-
2 x '-........:.~ :-::-:. 2 "3:::c:-!....---_-
I: 0.52 -.--..:;r.~.="= . -.......
2: 0.58
3: 0.65 6: 0.65 (annealed)
o 4: 0.72 5 . - .. -
5: 0.84 _ _
-1 "-"--
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

10

8 "'.

<S' 6 4.
3........ '"". '
....2~ "'-...
4 ::--.... ........
------~ ~.~: .
'"-=-.,....... ~ ..
2 -6------------ .......... _:.~~
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(b)

Figure 7. 7 Measured El and E2 spectra of Nix (Si02 )(1-x) granular films with
different Ni content.

According to Eqs. (7. 28) and (7. 29), the off-diagonal terms of the
complex dielectric tensor can be calculated from the measured magneto-
optical properties. Using the measured optical and magneto-optical constants,
the xy spectra can be calculated, as shown in Fig. 7. 8. Although the
measured magneto-optical Kerr spectra are proportional to xy' they will also
be significantly enhanced in the regions where optical constants have small
values, especially where 1 is close to 1 and 2 is small.
With this discussion in mind, it appears that the broad peaks in the Kerr
spectra are due to the combined effects of xy' which has relatively larger
values in the visible range with increasing x, and from the reduced optical
260 Liangyao Chen et a!.

0.02

o
---------~~
6 :;;.- ;::;;- _~
::. ~
-----
1~... <-- / 'x
<0~. -002
. 2... ---
"'3'--- -
4 ,~ I: 0.52
-0.04 /" 2: 0.58
/' 3: 0.65
.' 4: 0.72
-0.06 5: 0.84
6: 0.65 (annealed)
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

0.020 ...........
. ' ....-=-:-_.~.
<"~,
5 ,
~
.....e::!
,........
-........
..... ........
". ... ~~
"
X' ,
"
. '
........ 2.3'-.
~

"
,
',
" ~;;': ."-
1 '- ..... "-

/ 6 ~.

" '- .................. _-------


-0.0 I0 '--_...l..-_--'--_----1._ _.l...-_-'--_---'-
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(b)

Figure 7, 8 Calculated E xy spectra of Nix (Si~ ) 1- x granular films with


different Ni content.

constants that occur when x increases.

7.5.2.2 Co-Ag Granular Films


The magneto-optical Kerr effects of magnetic granular films (Wang et al. ,
2000a, 2000b), such as Co x Ag 1- x and Fe x Ag 1- x ' were studied in some detail
after the discovery of the large magneto-resistance of the materials. It was
found that for the Ag-based magnetic granular films, the Kerr effects were
enhanced after annealing at different temperatures. Figures 7.9 and 7. 10 show
the measured spectra of the complex dielectric function, and the Kerr rotation
and ellipticity, respectively, for C06 Ag 94 It can be seen that, before
annealing, the values of both 6k and Ek are very small and nearly zero over the
entire spectral range. When the annealing temperature is increased, the
absolute values of 6k and Ek increase, and 6k has a resonant peak in the 3.7-
4. 1 eV photon energy range. But the Ek spectrum has only a negative sharp
peak near 3. 85 eV. The peaks are shifted a bit towards lower energy with
increasing annealing temperature. This peak energy is close to the Ag plasma
frequency.
It is clear that there is a correlation between the Kerr effect and the
optical constants that are affected dramatically by annealing at different
Magneto-Optical Properties of Nanostructured Media 261

o
-5
- - as-dep.
--- 100'C
----- 250'C
-20
--_.- 400'C
-25 -.._ ..- 500'C
-30 L-_-'-_-'-_--'-_-----.J'-----_..L-_-'-
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

2.0 2.5 3.0 3.5 4.0 4.5


E(eV)
(b)

Figure 7. 9 The complex dielectric function spectra for Cas AQg4 annealed at
different temperatures.

0.4
- - as-dep.
0.3 - - - 100'C

0.2
~:~~~~~ J:~'\"
0.1

0
_ . - 500'C
~==-==.::=~.~~-. \
". _
~.~

-0.1 'J
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

0.1

0
.~~
-:,-::-~~.ro
.J:
v
.:
-0.1

-0.2
.I
\.
-0.3

-0.4
~
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(b)

Figure 7.10 The Kerr effect spectra of Cas Ag94 annealed at different temperatures.
262 Liangyao Chen et al.

temperature. This was also observed in metallic bilayer thin films, where it
was found that the magneto-optical response was significantly enhanced if the
adjacent nonmagnetic metallic layer had a low value of the optical constants in
the same wavelength range as the plasma frequency in the magnetic layer
CReim and Weller, 1989; Chen et aI., 1990; Katayama et aI., 1988). The
same mechanism may occur in metallic granular films, such as Co-Ag and Fe-
Ag films. In order to explore the origin of the magneto-optical response in the
granular films, the off-diagonal terms of the dielectric tensor and A / CA 2 +
B 2 ) , B/ CA 2 + B 2 ) were calculated, and are shown in Figs. 7. 11 and 7. 12.
0
~ I
-0.5 'I.
1!!1
.. /
-as-dep. ~~ ~l
---
~
-1.0 ---100"C ~~!1
,r!'
...l
N ----250"C '1"
~ -1.5 '.q
---400"C '1'
i1
'" -2.0
--500"C
~
!
-2.5
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

-2 '-;-----,;;'-;:----::!-:----;:l-;:------=-":-----,:-7--:-'-;
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(b)

Figure 7. 11 Calculated spectra of A / (A 2 + 8 2 ) and 8/( A 2 +82) for C06 AQg4 granular
films at different annealing temperatures.

In the spectral range covered, n, k and 2 are positive, and 1 changes


its sign depending on the dispersion of the spectra. The sharp peaks in 8k
occur when there is a minimum in CN + B 2 ). From Eq. C7. 28), this can occur
at the position where , ::::::::: 1 and 2 is very small, resulting in CA 2 + B 2 ) :::::::::
~ I I. For annealed C06 Ag 94 , this condition occurs at about 3. 85 eV, an
energy that is close to but slightly higher than the plasma energy and results in
the sharp peaks of 8k and k observed in the range.
Magneto-Optical Properties of Nanostructured Media 263

0.00

_ _ as-dep.

- - - 100L:
......... 250L:

- - - 400L:
_ .. - 500L:
-0.10
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E(eV)
(a)

0.00 .......-:_=_=_=_=---:=_~=--~
"""'------.~~:::.::::.. ~.~".
-0.Q2 ... ;?""
: -0.04 ....; ..
.
-0.06

- 0.08 7----::'-::---::-'-:----::-'-::-----="-=-----:-'-:--:-'-:
1.5 2.0 2.5 3.0 3.5 4.0 4.5
E (eV)
(b)

Figure 7. 12 Calculated spectra of xy for COl; Agg, .

To provide more detail, A/CA 2 + B 2 ) is negative over the entire spectral


range, but B / CA2 + B 2 ) changes its sign from negative to positive at the
position where E, = 1. Therefore, the negative peaks of 9k and Ek near
3. 85 eV are mainly due to the contributions of A / CA 2 + B 2 ) and E lxy at the
position where El=l and E2xy=O. The positive peak in 9k is due to CAE1XY+
BE2xy)/CA2+ B 2 ), but a cancellation in the term CAE2xy-BElxy)/CA 2 + B 2 )
gives a near-zero contribution to Ek at the higher photon energies, where the
onset of interband transitions occurs and results in higher values of E,xy and
E2xy' but with an opposite sign in the range.
Therefore, the enhancement of the Kerr effect will not only depend on the
optical constant changes with the annealing process, but also on the magnetic
structure in the granular Co-Ag films as presented by the function of Exy .

7.5.3 Magnetic Co-Cn Mnltilayers


Much attention has been paid to the importance of the optical and magneto-
optical properties of multilayered materials with sub-layers of nanoscale
thickness, since they show unusual properties not normally seen in the bulk
264 Liangyao Chen et al.

materials. For example, in multilayered combinations of ferromagnetic


transition metal layers with non-magnetic noble metal layers, one can
observed spin polarization of the non-magnetic component, e. g. , Pt in Co/Pt
multilayers (Samant et a!., 1994; Pizzini et a!., 1995; Atkinson and
Hendren, 1996a), thickness-dependent magneto-optical Kerr effect (MOKE)
oscillations (Zheng et al., 1998), ferro- and anti-ferromagnetic coupling
(Parkin et aI., 1990), and giant magnetoresistance (Baibich et a!., 1988,
1989 ). The optical and magneto-optical properties of some multilayer
structures are of direct interest. For example, Co/Pt, Co/Ni-Pt (Atkinson
et al., 1996b), and Co/Au (Atkinson et al., 1994) multilayers are being
considered as materials to be used in next generation magneto-optic
information storage systems. For other multilayered structures, such as
Co/Cu, the optical and magneto-optical properties are of indirect interest, as
they provide information about the applicability of optical multilayer theory to
understand the connections between electronic structure and optical properties
in ultra-thin metallic films.
In their pioneering work, Katayama et al. showed that the Kerr rotation in
Fe/Cu multilayers is enhanced at the absorption edge of Cu (1988). Since
then, extensive investigations of MOKE spectra of Fe, Co/noble-metal
bilayers and multilayers have been carried out (Xu et al., 1993a, 1993b,
1995; Sakurai and Shinjo 1993; Uba et al. , 1997; Atkinson and Dodd, 1997).
The observed features of the MOKE spectra are found to occur at the plasma
edge in the noble metal and are connected with magneto-optically-active
transitions in the noble metal as a consequence of the spin polarization of the
noble metal due to the proximity of magnetic layers at the interface.
In 1997, Uba et a!. (1997) carried out an experimental and theoretical
investigation of series of Co/Cu multilayers. Their main results are shown in
Fig. 7.13 (because of a difference in experimental conditions, their measured
Kerr rotations and ellipticities have signs that are opposite to conventional
signs). The samples were prepared using a "face-to-face" dc sputtering
system. The essential points of the experimental results can be summarized
as:
(1) The Kerr rotation spectra of the (CO/CU)40 multilayers show double
negative peaks and are much weaker than those of a pure Co thick film.
(2) As the amount of Co in the multilayers decreases, the spectra
generally become weaker, while there is an increase in the prominence of the
Kerr rotation peak located at about 2. 1 eV photon energy.
(3) In the energy range above 2. 1 eV, the Kerr ellipticity spectra for all
samples are quite similar. The negative peak position shifts to lower energy as
compared to the thick Co film. The zero crossing for the ellipticity spectra
occurs at about 2. 1 eV and has shifted from that for bulk cobalt at about
1.5 eV.
(4) The position of the broad peak just below 4 eV, corresponding to that
of cobalt, is independent of the Cu sublayer thickness. The negative peak at
the energy of 2. 1 eV is not observed in the pure Co film, and the amplitude
Magneto-Optical Properties of Nanostructured Media 265

0.1

Co/Cu
-1.49/1.01
-1.52/1.52
0.0 -......-1.51/2.17
_ _ 2.23/2.37
--Cobalt/2

-0.1
~

c:
0
"E0
e:::
-0.2

-0.3 ':- --':- ----' ...L.

0 2 4 6
E(eV)
(a)

0.2

0.1

0.0

o
:g -0.1
g
ill

-0.2

-0.3 0~--'---:-2-----'3'---...L- -. L - - . . . L .
4 5 6
E(eV)
(b)

Figure 7. 13 Measured values of the normal incidence polar Kerr rotation and ellipticity of
a series of Co/Cu multilayers.

and width of this peak changes with the Cu sublayer thickness. For the thinnest
Cu layer, the peak is transformed into a broad shallow minimum located in the
266 Liangyao Chen et al.

1. 5 - 2. 1 eV range. Obviously, the negative peak at about 1.5 eV is


attributed to the Co layers.
(5) As the amount of Co in the multilayers decreases, the spectra
generally become weaker, while there is an increase in the prominence of the
Kerr rotation peak located at about 2. 1 eV photon energy.
(6) In the energy range above 2. 1 eV, the Kerr ellipticity spectra for all
samples are quite similar. The negative peak position shifts to lower energy as
compared to the thick Co film. The zero crossing for the ellipticity spectra
occurs at about 2. 1 eV and has shifted from that for bulk cobalt at about
1.5 eV.
(7) It is generally clear that the complex MOKE spectra are dependent on
both the sub-layer thickness and the ratio of the Co/Cu sub-layer thickness.
Uba et a!. (1997) attempted to explain these results using first-principles
LMTO band-structure calculations for the Co/Cu lattice. They assumed an fcc-
( 111) texture and a lattice constant a = O. 3574 nm as an average of fcc
copper and cobalt lattice constants. Their calculations reproduced the main
broad features seen in Fig. 7. 13, but there were many unresolved questions
regarding the magneto-optical spectra. Especially, the calculated results
cannot explain the peaks that occur at the lower energy positions. Atkinson
and Dodd (1997) used effective medium theory to explain this data with some
success. In their work, they used the modified optical constant of the copper
layer as a result of increased scattering at the interface. This effect is taken
into account for the sub-layer thickness-dependent constants of the copper
layers through a modified relaxation-time term in semi-classical dispersion
theory for metals.
Starting with the macroscopic electrodynamics of multilayers and treating
the multilayer film as a single equivalent layer (SEL), the effective refractive
index n e and the effective magneto-optic parameter of the SEL Q e are given
by the relationships:
(7.64)
and
(7.65)
where
(7.66)
In Eqs. (7. 64) ...., (7. 66) the subscripts 1 and 2 refer to the sub-layer
media, for example Co and Cu, respectively.
For the thin films, the optical constants of metallic particles like Co and
Cu not only depend on their size, but also depend on the film thickness. Thus,
the dispersion relation for copper can be expressed by
fw 2
E(W) = 1 + aEB - w(w + tL/RT) (7.67)

where w p is the plasma frequency, f is the oscillator strength, L is the mean


Magneto-Optical Properties of Nanostructured Media 267

free path, R is the radius of the particles, and T is the relaxation time
associated with the Drude free-electron theory. The second term aEe is the
contribution to E (w) from the interband transitions of the bound electrons at

0.1

Co/Cu
- - - 1.49/1.01
0.0 - - 1.52/1.52
---1.51/2.17
\ ........ 2.23/2.37
~

J
~

c::
-0.1
"\
\ ':.;\
\ :
\'\
\\ :

-0.2 \\\ l/
\
\\ \~ . ---:,
/' .: I

\ .
\ 00/
,,-,'\ I
\J '\ /
-0.3 l - -'- ...._---L- --'c-
0 2 4 6
E(eV)
(a)

0.2

0.\

0.0
~

o
:~
]- -0.\
Ul

-0.2

-0.3'-----......L..----'--------'--
o 2 4 6
E(eV)
(b)

Figure 7.14 Calculated values of the dispersion of the polar Kerr rotation and ellipticity for
Co/Cu multilayers based on the bulk optical and magneto-optical constants of cobalt and
modified optical constants for ultra-thin copper.
268 Liangyao Chen et al.

the bulk and ultra-thin film interface.


Using the measured optical constants of Co and the magneto-optical
constant of Co deduced from the measurements of the Kerr effect, the Kerr
effects of Co/Cu multilayers can be calculated, and the resulting curves are
illustrated in Fig. 7. 14. The calculated and measured results are in good
agreement with each other.

7.6 Summary

We have presented a necessarily brief and incomplete summary of magneto-


optical effects in nanostructured magnetic materials, and related the results on
nanostructured systems to those obtained in pure bulk materials. We have
chosen a few examples to illustrate the interesting and unusual magneto-optical
phenomena that have been found in films with granular or multilayered
structures. It is clear that such systems will be widely studied in the future both
because of the potential for new physics and because of their great potential
for new applications to modern technologies. Measurements of magneto-
optical properties will continue to be a powerful and useful tool as the research
in this area unfolds.

References
Argyres P. N. Phys. Rev. 97: 334(1955)
Atkinson R., W. R. Hendren. J. Magn. Soc. Jpn. 20: 291C1996a)
Atkinson R. , S. Pahirathan, I. W. Salter, P. J. Grundy, C. J. Tatnall, J.
C. Lodder and Q. Meng. J. Magn. Magn. Mater., 162: 131C1996b)
Atkinson R., W. R. Hendren, I. W. Salter and M. J. Walker. J. Magn.
Magn. Mater. 130: 442 ( 1994)
Atkinson R. and P. M. Dodd. J. Magn. Magn. Mater. 173: 202(1997)
Baibich M. N., J. M. Broto, A. Fert, F. V. D. Nguyen, F. Petroff, P.
Etienne, G. Creuzet, A. Friederich, and J. Chazelas. Phys. Rev. Lett.
61: 2472(1988)
Baibich G., P. Grunberg, F. Saurenbach and W. Zinno Phys. Rev. B 39:
4828(1989)
Bennet H. S. and E.A. Stern. Phys. Rev. 137: 448(1965)
Berkowitz A. E., J. R. Mitchell, M. J. Carey, A. P. Young, S. Zhang, F.
E. Spada, F. T. Parker, A. Hutten and G. Thomas. Phys. Rev. Lett. 68:
3745(1992)
Chen L. Y., W. A. McGahan, Z. S. Shan, D. J. Sellmyer and J. A.
Woollam. J. Appl. Phys. 67: 7547(1990)
Chen L. Y., X. W. Feng, Y. Su, H. Z. Ma and Y. H. Qian. Appl. Opt. 33:
1299(1994)
Magneto-Optical Properties of Nanostructured Media 269

Chen L. Y., S. M. Zhou, Y. Su, H. Z. Ma, P. J. Zheng, Y. H. Qian,C. J.


Chen, X. Z. Wang and W. Giriat. Solid State Commun. 98: 829(1996)
ChenL. Y., S.M. Zhou, Y.X. Zheng, Y.Cheng, Y. M.Yang, Y. H. Qian,
C. H. Shang and Y.J. Wang. Optical Engineering 36: 3188(1997)
Deng S. H., S. Y. Wang, J. Li, Z. Liu, Y. L. Chen, Y. M. Yang, L. Y.
Chen, H. Liu, X. X. Zhang and D. Lynch. Acta Physica Sinica 50: 169
(2001) .
Ebert H. and H. Akai. J. Appl. Phys. 67: 4798 ( 1990)
Erskine J. L. and E.A. Stern. Phys. Rev. B 8: 1239(1973)
Faraday M. Trans. Roy. Soc. (London) 5: 592( 1846)
Feil H. and C. Haas. Phys. Rev. Lett. 58: 65(1987)
Hulme H. R. Proc. Roy. Soc. A 135: 237(1932)
Katayama T. and K. Hasegawa. In: Proc. 4th International Conf. on Rapidly
Quenched Metals, Sendai, p. 915 (1981)
Katayama T., Y. Suzuki, H. Awano, Y. Nishihara and N. Koshizuka. Phys.
Rev. Lett. 60: 1426(1988)
Kerr J. Philos. Mag. 3: 339(1877)
Kerr J. Philos. Mag. 5: 161(1878)
Kittel C. Phys. Rev. 83: 208(951)
Manens J. W. D., W. L. Peeters and P. Q. J. Nederpel. J. Appl. Phys.
55: 1100 ( 1984)
Mansuripur M., F. Zhou and J. K. Erwin. Appl. Opt. 29: 1308(1990)
Parkin S. S. P., N. More and K. P. Roche. Phys. Rev. Lett. 64: 2304
(1990)
Pizzini S., A. Fontaine, C. Giorgetti, E. Dartyge, J. F. Bobo, M. Piecuch
and F. Baudelet. Phys. Rev. Lett. 74: 1470(1995)
Reim W., O. E. Huesser, J. Schoenes, E. Kaldis, P. Wachter and K.
Seiler. J. Appl. Phys. 55: 2155(1984)
Reim W. and D. Weller. IEEE Trans. Mag. 25: 3752( 1989)
Reim W. and J. Schoenes. In: K. H. J. Buschow and E. P. Wohlfarth, eds.
In: Magneto-Optical Spectroscopy of f-Electron Systems. Elsevier Science
Publishers B. V. (1990)
Sakurai M. and T. Shinjo. J. Appl. Phys. 74: 6840(1993)
Samant M. G. , J. Stohr, S. S. P. Parkin, G. A. Held, B. D. Hermsmeier,
F. Herman, M. Van Schilfgaarde, L. C. Duda, D. C. Mancini, N.
Wassdahl and R. Nakajima. Phys. Rev. Lett. 72: 1112(1994)
Sato K. Jpn. J. Appl. Phys. 20: 2403(1981)
Sato K. , Y. Aman and H. Hongu. J. Magn. Magn. Mater. 104: 1947(1992)
Singh M., C.S. Wang and J. Callaway. Phys. Rev. B 11: 287(1975)
Suits J. C. Rev. Sci. Instrum. 42: 19(1971)
Takanashi K., K. Sato, J. Watanabe, Y. Sato and H. Fujimori. Jpn. J.
Appl. Phys. 30: 52(1991)
Uba S., L. Uba, A. Y. Perlov, A. N. Yaresko, V. N. Antonov and R.
Gontarz. J. Phys. Condo Matt. 9: 447(1997)
Uba L. , S. Uba, V. N. Antonov, A. N. Yaresko and R. Gontarz. Phys. Rev.
270 Liangyao Chen et al.

B 62: 16,510(2000)
Wang C. Sand J. Callaway. Phys. Rev. B 9: 4897(1974)
Wang J. Q., P. Xiong and G. Xiao, Phys. Rev. B 47: 8341(1993)
Wang S. Y. , L. Y. Chen, W. M. Zheng, R. J. Zhang, D. L. Qian, Y. X.
Zheng, X. F. Jin and B. Y. Li. J. Infrared Millim. W. 18: 243(1999)
Wang S. Y. , Z. C. Shen, W. M. Zheng, Y. X. Zheng, S. M. Zhou and L. Y.
Chen, Physica B 279: 109(2000a)
Wang S. Y., H. Y. Li, S. M. Zhou, J. Li, and L. Y.Chen. SPIE, 4086: 352
(2000b)
Xiao G. and C. L. Chien. Appl. Phys. Lett. 51, 1280(1987)
Xiao J. Q., J. S. Jiang and C. L. Chien. Phys. Rev. Lett. 68: 3749(1992)
Xiao G. , J. Q. Wang and P. Xiong. Appl. Phys. Lett. 62: 420(1993)
Xiao G. and J. Q. Wang. J. Appl. Phys. 75: 6604(1994)
Xiong P., G. Xiao, J. Q. Wang, J. Q. Xiao, J. S. Jiang and C. L. Chien,
Phys. Rev. Lett. 69: 3220(1992)
Xu Y. B., Q. Y. Jin, Y. Zhai, M. Lu, Y. Z. Miao, Q. S. Bie and H. R.
Zhai, J. Appl. Phys. 74: 3470( 1993a)
Xu Y. B., Q. Y. Jin, Y. Zhai, Y. Z. Miao, M. Lu and H. R. Zhai. J. Magn.
Magn. Mater. 126: 541(1993b)
Xu Y. B., M. Lu, B. S. Bie, Y. Zhai, Q. Y. Jin, X. B. Zhu and H. R.
Zhai. J. Magn. Magn. Mater. 140: 581(1995)
Yaresko A. N., L. Uba, S. Uba, A. Y. Perlov, R. Gontarz and V. N.
Antonov. Phys. Rev. B 58: 7648(1998)
Zhang R. J., H. Y. Li, X. S. Zhu, P. Zhou, J. Li, S. Y. Wang, Z. C.
Shen, Y. X. Zheng and L. Y. Chen. J. Infrared Millim. W. 20: 57(2001)
Zheng W. M., S. Y. Wang, Q. L. Qian, Y. X. Zheng, S. M. Zhou, L. Y.
Chen, X. L. Shen and H. Xia. Phys. Status. Solidi. A 167: 223(1998)
8 Magnetoresistive Recording Heads

Christopher D. Keener

8. 1 Introduction

Magnetic data storage is a crucial contributor to today's economy and will


continue to be relied upon in the foreseeable future. Far from being restricted to
home computer use, magnetic storage enables efficiency in a vast array of
industries such as banking, telecommunications, all of engineering and
science, and entertainment, and storage fuels the explosive growth of the
internet. Hard disk drives alone account for a $ 25 billion revenue annually,
with over 175 million units shipped in 1999 (Monroe, 2000), not including
high density magnetic tape drives and other applications using removable
media.
One reason that magnetic recording is playing such a central role today is
the stunning progress in technology, and the most complex component of the
hard disk drive (HDD) is the magnetic recording head. Areal density is
defined as the number of bits that are stored per unit of disk surface area.
When the HDD was invented by IBM in 1957, the areal density was 2 kbits/in2
and the data rate was 12.5 kB/s (Ashar, 1997, p.3). Since then, areal
density has grown, at least exponentially. Largely because of sensors using
the giant magnetoresistance effect, in the last 3 years areal density has
doubled each year, exceeding even the growth rate of transistor density in the
semiconductor industry. Recently, mobile HDD products with 26 Gb/in2 , or
15 GB of data on a single 2. 5" disk have begun to ship (IBM Travelstar 30GN,
2001 ), as well as server products operating at data rates exceeding 650 Mb/s
(IBM Ultrastar 73LZX, 2000). Figure 8. 1 charts the remarkable advance of
HDD areal density over the last 44 years. Although lack of market demand for
higher-capacity disk drives may eventually slow the pace of development, the
availability of high-capacity storage at low costs is already spawning new
markets for items such as the personal video recorder, which can pause live
TV, skip the commercials, and store up to 35 h of programming on a 40 GB
drive (Langberg, 2001).
To understand the function of recording heads, it is helpful to survey the
entire magnetic recording process. Data is written to a disk in concentric
circles called tracks, using an inductive write head (Ashar, 1997, chapter 5) .
272 Christopher D. Keener

Areal density perspective


45 years of technology progress

~";' 10
4

~ 10 3
13
~ 10
2

E I st thin film head


.~ 10 c
c<l.l
-0
;3
<e
<l.l IBM disk drive products ~
~ 10-1 Industry lab demos I

10- 2
IBM RAMAC (First hard disk drive)
10-3 L - -
19-'-6-0---19-'-7-0---1....l98-0---1....l9-90---2-'0-00---2-'010
Production year

Figure 8.1 Growth in HOD areal density over time.

Tracks at a single radius on multiple disk surfaces are called a cylinder. The
areal density is the track density, or number of tracks per inch, multiplied by
the linear density, or number of bits per inch within a track. While the disk is
spinning at rotational speeds up to 15, 000 r/min, a voice coil motor (VCM)
consisting of a magnet and Cu coils forces the actuator to rotate, bringing the
head to the desired track. The head stays on track via a servo system in which
some of the data written on the track contain the track position information,
and the position error signal feeds back to control the current in the VCM.
Figure 8. 2 illustrates how the head reads the signal from the disk. Current
technology uses magnetoresistance to detect fringing fields from the media.
Bits are recorded where the magnetization reverses at "transitions." These
back-to-back magnets produce fields with a component emerging from the
plane of the disk. Transitions alternate in polarity, so that the resulting
waveform of head output is a series of alternating positive and negative
pulses. The head "flies" over the disk on a slim cushion of air at a height as
low as 10 nm. With an electrical current in the sensor, when the resistance
changes in response to the magnetic field, a voltage is generated that is
amplified and passed to a read channel. The channel extracts digital data from
the noise using complex algorithms (known as partial response maximum
likelihood or partial response maximum likelihood (PRML in which the most
probable combination of a series of bits is determined by minimizing the
difference of sampled data from ideal target values that would be observed in
the absence of noise. The written data contains some redundancy that allows
the channel to check whole blocks of data for errors, and to make corrections
on the fly or to retry reading on the next revolution of the disk. The entire HOD
is controlled by a programmed microprocessor with much of the firmware
stored on a dedicated portion of the disk itself.
Magnetoresistive Recording Heads 273

MRW

Magnetic spacing
~--_ _ I = FH+DOC+HOC+rec

Figure 8.2 Schematic of the readback process. The black arrows in the media refer to
its magnetization direction. Arrows emerging from the media refer to the fringing fields at
the transitions. which produce the signal in the GMR sensor. The large gray arrow labeled
v refers to the direction of motion of the media as the disk spins under the head.

The primary limiting factors for areal density are the signal-to-noise ratio
(SNR) and pulse width (PW50). Magnetic fields from adjacent transitions
oppose one another; this phenomenon is referred to as inter-symbol
interference (lSI). As PW50 decreases, lSI diminishes. However, at high
densities, lSI may become so severe that the channel is not able to determine
the correct bit sequence in the midst of the noise, and errors occur. A raw bit
error rate (SER, soft error rate) of about 1 error per million bits exceeds the
channel's error correction capability, and performance begins to degrade
sharply as more errors occur.
Whereas the linear bit density mainly is limited by inter-symbol
interference, track density is limited by SNR. Readback signal amplitude is
proportional to the magnetic read width of the sensor. Thus, more amplitude
per unit trackwidth is needed in order to increase storage capacity. In the
early 1990' s, IBM adopted magnetoresistive heads for HODs because the
signal amplitude is higher than that of inductive read heads. These heads
utilized the anisotropic magnetoresistance of permalloy (Ni so Fezo) , about
O. 5 %, to sense the applied field from the disk. However, the giant
magnetoresistance (GMR) effect was discovered in 1988 (Baibich et al. ,
1988). The change of resistance in a small applied magnetic field can be over
20 %, and the effect was quickly harnessed for a practical device called a spin
valve (Dieny et al., 1991). In 1998, just 10 years after the discovery of
GMR, IBM introduced the first GMR head into a product (IBM Travelstar 6GT,
1998), and other companies soon followed. GMR heads have completely
eclipsed the anisotropic magnetoresistance (AMR) heads in current HOD
products.
274 Christopher D. Keener

8. 2 Physics of Magnetoresistance

The first magnetoresistive recording heads utilized AMR effect, and this type
of head is still used in many technologies today such as tape heads. Figure 8.3
illustrates the three layers of an AMR head. The first layer is known as a soft
adjacent layer, or soft adjacement layer (SAL), because of its soft magnetic
properties. Electrical current flows through the sensing element, or stripe,
between the leads that define the magnetic read width (MRW). Ampere's
Law dictates that a magnetic field surrounds the current flow as shown in the
figure. The SAL is saturated in the transverse direction, perpendicular to the
MRW, biased by the current-induced magnetic field. The second layer is a
poorly conducting non-magnetic spacer material. The third layer is the active
AMR layer, generally made of permalloy, which was chosen for its relatively
high AMR, low magnetostriction, and chemical stability. Ideally the
orientation of the sensing layer in the absence of externally applied magnetic
field is at 458 from the transverse direction, and biasing of the sensing layer is
provided by a combination of the fringing field (also known as demagnetizing
field) from the SAL and the current-induced field. The magnetic field from the
media causes the magnetization of the AMR layer to rotate, producing a signal
via the AMR effect.
SAL

Ta spacer

'-- ....... / 0
7 ,. I_bias
H
Figure 8. 3 Diagram of the AMR head structure.

Anisotropic magnetoresistance is defined to be the difference in electrical


resistivity with current flowing parallel to the magnetization from that with
current flowing perpendicular to the magnetization, expressed as a fraction of
the total resistivity:

AMR = P II - P1-
P II
where PII is electrical resistivity with current flowing parallel to the
magnetization (in AMR material) .
Magnetoresistive Recording Heads 275

The AMR effect arises from the spin-orbit coupling of the conduction electrons
(Campbell et aI., 1970, 1980; Jaoul et aI., 1997). The sensor in Fig. 8.3 has an
output signal proportional to AMR coff f3, where f3 is the deviation of the NiFe
layer magnetization from the longitudinal direction (Ashar, 1997, p. 125).
Current HOD technology uses the GMR effect. The GMR effect results
from spin-dependent scattering that occurs predominantly at the interface
between ferromagnetic and non-magnetic layers of a multilayer material
(Nesbet, 1998). Conduction electrons are polarized so that the majority of
their spins are parallel to the magnetization (call this direction m). Because of
this spin polarization, most of the electronic dm states are occupied. Most
conduction electrons are in the low-effective-mass sand p states. Dissipative
scattering involves a change of energy (tor example, when s or p electrons are
scattered into d states). Neglecting spin flip scattering, more scattering of so
and po electrons into do states can occur than sm and pm into dm states,
which are unavailable. As a result of this preferential scattering, the resistivity
parallel to the magnetization, pm is lower than po. Resistivities pm and po
can be thought of as parallel resistance paths, and the total resistivity is
described by the parallel resistor model as p = pm// po = pmpo/ (pm + po).
When the magnetization of the ferromagnetic layers is aligned, the spin m
carriers have a lower' resistivity in all the ferromagnetic layers. When alternate
layers are antiparallel , spin m electrons in one layer are spin 0 in the next
layer. Thus there is no low-resistivity path for either spin type through all of
the layers. So the resistance of the antiparallel state is higher. The difference
may be described quantitatively by the GMR coefficient,

GMR = 1i.R/R = Rap - R II (8.1)


R II

where 1i.R is the change in resistance of the head between the two states in a
two-state system (magnetic telegraph noise), 1i.R / R is GMR coefficient, or
change in resistance resulting from applying a large magnetic field, expressed
as a fraction of the resistance at zero field, R II is the resistance of the
multilayer with all of the ferromagnetic layers aligned in the same direction,
and R 8p is the resistance with every other layer reversed. The plot of
resistance vs. magnetic field is known as a transfer curve. The idealized
transfer curve of the GMR head is
R = R o + (1i.R/R)sin8 (8.2)

where 8 is the angle of the free layer magnetization with respect to the
longitudinal direction (Tsang et al. , 1998).
In data storage applications, a specific multilayer device called a spin
valve, sketched in Fig. 8.4a, is used (Dieny et aI., 1991). A spin valve
contains a ferromagnetic layer that is pinned by an antiferromagnet, a
conducting spacer layer, and another ferromagnetic layer that freely rotates in
the applied magnetic field. Exchange coupling (Ruderman and Kittel, 1954;
276 Christopher D. Keener

Jones, 1998) between the antiferromagnet and the pinned layer produces a
unidirectional anisotropy in the pinned layer, keeping its magnetization
direction fixed even in the presence of an applied magnetic field from the disk.
Exchange coupling, also known as the Ruderman, Kittel, Kasuya, and
Yoshida (RKKY) interaction, is a short range oscillatory interaction. It arises
as polarized conduction electrons form a cloud around magnetic ions. These
nearby electrons screen the magnetic moment of the ion, so that conduction
electrons at a larger radius become polarized in the reverse direction. As a
result, the coupling of magnetic atoms as a function of their distance alternates
between ferromagnetic and antiferromagnetic with a period on the order of 3.
The RKKY interaction is very strong at short range, producing anisotropy fields
of the pinned layer up to several kOe. The antiferromagnetic material in a spin
valve device is chosen to achieve high pinning fields, and a high blocking
temperature is desirable so that pinning remains effective even at high
temperatures that may be reached inside an operating disk drive.
-H_D

NiFe
_H_'
Nanolayer

Pinned layer Pinned layer

(a)

-H_D

NiFe o -H_I
Free layer
Nanolayer
~~(fCf~~~~/. (!)
////;~////////////////////////////////////.
, bias
Pinned layer 2

Ru
Pinned layer I
Pinned layer

(b)

Figure 8.4 (a) spin valve structure; (b) structure of AP-pinned spin valve. Magnetization
directions of the layers are indicated, as well as demagnetization and bias current-induced
magnetic fields acting on the free layer.
Magnetoresistive Recording Heads 277

Because there is only one pair of interfaces at which differential scattering


can occur in a spin valve, great care is exercised in controlling the scattering
at these interfaces. First, in general, while NiFe is desirable in the free layer
for its soft magnetic properties, a dusting of a higher-moment material such as
Co, known as a nanolayer, is applied at the interface with the conducting
spacer to enhance the differential scattering (Kanai et aI., 1996). Second,
underlayers and capping materials (on top of the free layer) are designed
to reduce the amount of inelastic scattering at these boundaries where
scattering is not spin dependent (Egelhoff et al. , 1995, 1997). The goal is to
produce specular, reflective, or mirror-like scattering of electrons back
toward the interfaces of the magnetic layers with the conducting spacer.
Third, materials are engineered with careful regard for the microstructure,
interfacial roughness, and reliability of the final product. In this way, GMR
coefficients of up to 20 % have been demonstrated in spin valve devices (Koui
et al. ,2001).
Besides the GMR coefficient, a number of other factors playa significant
role in sensor amplitude. Further, SER in the disk drive is governed mainly by
SNR, or amplitude at high frequency divided by the rms of total integrated
noise power,

SNR = Amp (8.3)


Jt Sv(f)df

where f o ' attempt frequency in a thermally activated and two-state system, is


defined to be the maximum data rate, S v ( f) is the noise power spectral
density and Amp is the peak-to-peak amplitude of a magnetic sensor, where f
is frequency. This noise includes contributions from the disk (which are
amplified by a high-amplitude head), and Johnson noise from the head and
preamplifier (noise power = 4kTR tJ.f, where k is Boltzmann constant and tJ.f
is bandwidth). Since most of the disk noise is associated with the transitions
(Murdock, 1992), the noise measurement needs to be performed with a high-
frequency data pattern written to the media, rather than on DC-erased media.
When disk noise dominates, increased amplitude does little to improve SER,
but in today's hard drives, disk, head, and preamplifier noise are all
significant. If the sensor is biased at a constant voltage, amplitude is
approximately proportional to tJ.R / R, sensor sheet resistance, bias voltage,
MRW, excitation (fraction of the full transfer curve that is utilized by the
applied magnetic field coming from the media), and inversely proportional to
magnetic spacing. Other factors, such as the stripe height, hard bias
thickness, and intrinsic materials properties of the free layer, influence
amplitude.
Reliability considerations limit the maximum bias voltage that can be
used. Joule heating causes the temperature to increase sharply with increasing
bias current, as tJ. T ..... [2 R. Diffusion of atoms within the sensor can degrade
278 Christopher D. Keener

interfaces and reduce amplitude. Most processes that cause amplitude


degradation are thermally activated. Thus head lifetime, defined by the
amount of time that passes before a head loses a certain percentage of
amplitude (e.g., 20%), is proportional to eEo/kT, where Eo is an activation
energy and T is the temperature of the sensor. Most of the heat is conducted
away from the sensor to the shields across the insulating gaps. By choosing
gap materials with improved thermal conduction, bias voltage may be
increased, with direct benefit to head amplitude (Hayakawa, 2000). Besides
heat conduction into the shields, some heat is dissipated into the leads, and
thus the temperature of the sensor is actually not uniform (Granstrom and
Tabat, 2000).
One spin valve design with many advantages utilizes a synthetic
antiferromagnet for the pinned layer, or an AP-pinned spin valve structure
(Heim and Parkin, 1995; Beach et al. , 2000; Pinarbasi et al. , 2000). The
synthetic antiferromagnet, illustrated in Fig. 8.4b, consists of a thin non-
magnetic film such as Ru sandwiched between two ferromagnetic layers,
where the thickness of the non-magnetic layer is controlled to produce a strong
antiferromagnetic exchange coupling between the two ferromagnetic layers. A
benefit of this configuration is to reduce the total moment of the pinned layer,
decreasing the energy of the pinned layer so that its magnetization is more
thermally stable.
It is generally desirable to have a small asymmetry between positive and
negative pulse amplitudes. Asymmetry is a function of both the bias point and
excitation. The bias point is defined as the point on the transfer curve with no
externally applied magnetic field. Several internal magnetic fields must be
balanced in order to produce a bias point near the middle of the transfer curve,
as sketched in Figs. 8.4 and 8.5. Due to magnetostatic coupling between the
pinned layer and the free layer, in the absence of bias current the antiparallel
state has lower energy than the parallel state. This interaction can be
expressed as if the pinned layer were exerting a magnetic field, known as a
demagnetization field, on the free layer, H 0 = 8M r t / SH, where M r t is the
remanent magnetization of the pinned layer multiplied by its thickness, where
SH is stripe height. One advantage of the synthetic antiferromagnetic pinned
layer is that its lower net magnetic moment results in a lower demagnetization
field acting on the free layer, making the bias point easier to control. In
addition to the magnetostatic field, there is a weak exchange coupling
between the pinned layer and the free layer, because the nonmagnetic spacer
layer is so. thin (on the order of 20). This exchange field, known as the
ferromagnetic coupling field, generally needs to be kept near a minimum due
to its strong sensitivity to the spacer layer thickness. The bias current induces
a magnetic field opposing the demagnetization field, given by Ampere's law,
H,""" 1/ SH, where SH is tripe height. Finally, uniaxial magnetic anisotropy of
the free layer can be expressed as a magnetic field H k that influences both the
bias point and the amplitude.
Magnetoresistive Recording Heads 279

2r--------+----l--~

----s-
---------
Readback wavefonn

---- -- -- -----
-------- ------
Asym<O

-1

-2 2

Figure 8.5 Spin valve transfer curve, illustrating the effect of bias current on asymmetry.
Asymmetry increases toward zero as bias current is increased.

The externally applied field H is proportional to M, t, the remanant


magnetization of the disk multiplied by its thickness. As discussed previously,
to obtain high linear density, it is necessary to shield the sensor from the
magnetic fields of transitions that are not located directly underneath the
sensor. In other words, a pulse width of short duration is needed to distinguish
one pulse from the next. The pulse width at half maximum defines PW50,
which is given by

PW50 = .)(g/2)2 + 4(0 + d)(o + d + 0) (8.4)

where 9 is the total distance between the two shields, d is the magnetic
spacing (distance) between the free layer and the media surface, 0 is the
magnetic thickness of the media, and 0 is the width of the transition that is
written into the media (Wang and Taratorin, 1999, p.146; Bertram, 1994,
p. 133; Taratorin, 1996, p.l0; Ashar, 1997, p.57). Because much of the
media flux is diverted into the shields, low-frequency amplitude is roughly
proportional to gap thickness. If the gap thickness is nearly optimal, this loss
of low-frequency amplitude with decreasing gap thickness is offset by
improved resolution of the sensor at high linear density, so that high-frequency
amplitude and achievable areal density are relatively insensitive to gap
thickness. One consequence of the shields is that flux from the media leaks
across the gaps into the shields as it travels from the near edge of the sensor
toward the back of the stripe. Thus the field appl ied to the free layer, and the
resulting angle of rotation within the free layer away from the longitudinal
direction, decrease with distance from the near edge of the sensor (Potter,
1974).
280 Christopher D. Keener

8. 3 Magnetoresistive Head Fabrication

Thousands of heads are fabricated on a single Ab03/TiC substrate (Hsiao,


1999; Tsang et aI., 1998). Before the spin valve layers are deposited, the
first ferromagnetic shield and insulating gap are deposited. Typically, the spin
valve layers are sputtered, and then patterned and etched through a
photolithographic process to define the track width and the back of the stripe.
Thicknesses of individual layers in a spin valve must be controlled to sub-
atomic tolerances in order to maintain consistent performance among millions
of heads. Normally, a hard magnetic material is used at the ends of the sensor
to maintain the free layer in a single domain state (Kroumbi, 1991). Leads
are added and routed to large pads for later connection to the preamplifier.
The second insulating gap is deposited on top of the sensor and the second
ferromagnetic shield is electroplated (Fontana, 1995). Finally, the inductive
write head is fabricated on top of the second shield.
The finished wafer is cut into rows and subsequently into individual heads.
The head on its substrate is called a slider. The slider is carefully polished, or
lapped, to remove excess material from the sensor, resulting in a stripe height
that can be controlled to within a few 10 s of nm. An air bearing surface is
patterned on the lapped surface of the slider (on the surface that had formed
the boundary between rows on the wafer) (Chhabra et al., 1994). This air
bearing enables the head to hover over the spinning disk at a constant height
on a slim cushion of a few air molecules, which can be thinner than 10 nm. An
individual slider is mounted onto a suspension to form head gimbal assembly
(HGA) , and several HGAs are mounted together on a rotatable actuator with
an attached preamplifier, making up what is known as a head stack assembly
(HSA) , which is placed into the HDD.

8.4 Noise in Magnetoresistive Sensors

Ferromagnetic domain wall motion produces either random telegraph noise or


1/ f noise in ferromagnetic thin films (Keener and Weissman, 1995). This type
of behavior was observed in giant magnetoresistance (GMR) materials
(Hardner et al. , 1995; Kirchenbaum et aI., 1995) and in spin valve sensors
(Zhang et al., 2001). In spin valves, metastable states of the free layer
Magnetoresistive Recording Heads 281

magnetization give rise to random telegraph noise, baseline popping noise,


and bimodal amplitude in the HDD, as well as kinked or open-loop transfer
curves. These symptoms are referred to as magnetic instability and are
illustrated in the waveforms and transfer curve shown in Fig. 8. 6. Random
telegraph noise is thermally activated switching between two nearly
degenerate states of the magnetization in a portion of the sensor, resulting in
switching of the sensor resistance via the GMR effect. Switching noise can be
seen by the channel as data and thereby degrade the SER. Furthermore,
baseline popping instability often causes the servo system to mis-identify the
cylinder, making it difficult for the head to locate data or remain on a single
track. To reduce the frequency of occurrence of instability in heads, at the
ends of the sensor, a hard bias layer is used to exert a magnetostatic field to
pin its magnetic orientation longitudinally (along MRW) (Kroumbi, 1991).
The magnetization of the free layer is coupled to the sensor resistance via
the GMR effect. Based on Eq. (8. 2), the amplitude of a random telegraph
noise signal is given by

.6.R = (AR/R) V o.6.(sineo) (8.5)


VFL
where V0 is the volume of the region that is unstable, A (sine o ) is the
difference in sine between the two states, eo is the deviation of the unstable
region's magnetization from the longitudinal direction, and VFL is the volume of
the free layer. The fluctuations are thermally activated, and thus the duty
cycle obeys an Arrhenius relationship (Keener and Weissman, 1995),
tu / td = const x eAMzHlkT (8.6)
where .6.M z is the transverse component of the difference in magnetic moment
between the two states, MA (sine). Because transfer curves are typically
measured at low frequency, when the magnetic field is tuned to make the
energies of the two states similar, rapid telegraph noise is time-averaged to
produce a smooth kink in the transfer curve. The width of the transfer curve
kink is governed by the duty cycle given in Eq. (8.6), so that

AH = kTln81 (8.7)
AM z
From Eq. (8.5) we see that significant instability may be caused by unstable
regions of smaller volume as the free layer dimensions are reduced in order to
increase areal density. From Eq. (8. 7), we see the consequence that the
typical kink width is becoming broader. Thus instability is becoming more
difficult to detect in transfer curve measurements.
Magnetic instability in spin valves is a challenging engineering problem
because most design changes that improve the sensor amplitude also
aggravate instability. For example, while free layer thickness reduction
282 Christopher D. Keener

(a)

y
(b)

(c)

Figure 8. 6 (a) example of a waveform from a head that exhibited random telegraph
noise; (b)example of baseline popping noise, defined as random popping of the baseline
between different levels, where the popping is associated with the pulses in the
waveform. This head also exhibited some telegraph noise (head switching between states
without requiring a transition from the disk to excite it) - e. g., at 1. 2 divisions;
(c) example of bimodal amplitude-large variability of positive-pulse amplitude, whereas
the negative pulse amplitude remains relatively constant; (d) transfer curve of a head with
a kink. This head also exhibited telegraph noise. Continued on page 283.
Magnetoresistive Recording Heads 283

600

400

i I

-100 -50 50 100


v (IlV)

-600

-800

-1000
H(Oe)
(d)

Figure 8. 6 Continued from page 282.

increases the excitation of the sensor (utilization of more of the transfer


curve), it nearly always increases the number of heads detected with
instability (Zhang et al., 2001). Increasing hard bias thickness improves
stability but stiffens the free layer and reduces the active magnetic read width
of the sensor, costing amplitude.

8. 5 Future of Magnetic Data Storage: Challenge in


Maintaining Rapid Pace of Development

So far, the maximum GMR coefficient demonstrated in spin valves with


reasonably low ferromagnetic coupling field and free layer coercivity (Koui
et al. , 2001) is <20%. Head sensitivity at narrow MRW may be improved
with alternate free layer stabilization methods such as exchange tab (Zhu et
al. ,2001). Using alternative gap materials may permit higher bias voltage due
to improved thermal conduction from the sensor, resulting in improved
amplitude (Hayakawa, 2000). Beyond these incremental improvements, it
appears that most of the advantages of spin valves are already being realized
in today's products.
Besides the read head, there are myriad problems with other HDD
components. For example, the challenge of the servo keeping the head
within < 100 nm of the track center may require a dual stage actuator
284 Christopher D. Keener

(Mcinerney, 2000). The superparamagnetic effect, spontaneous, thermally


activated magnetization reversal of small grains in the media (analogous to
magnetization reversal that causes telegraph noise in heads), limits the ability
to reduce grain size to improve media noise. Even with a reduction in the net
magnetic moment of the media using antiferromagnetically-coupled media
(analogous to the synthetic antiferromagnet used for the pinned layer in
heads), it is becoming increasingly difficult to achieve higher densities. To
maintain the thermal stability of the media, high coercivity is needed, and it is
difficult to find useful high-moment materials to write on such high-coercivity
disks. Channels are approaching their theoretical limit; above a certain limit
of PW50/To , where To is the bit period, there is insufficient information
available to reconstruct the bit sequence. Given all of these difficulties with the
head as well as other HDD components, it seems likely that the areal density
growth rate will slow down somewhat after the next 1- 2 years.
Tunnel valves have been reported in publications as potential candidates
for future magnetic recording heads (Gangopadhyay, 2000). Thus far,
however, whenever tunneling magnetoresistance is high, resistance is also
high. Johnson noise is too large to obtain an SNR that permits acceptable error
rates. If spin valves had not already become the industry standard, tunnel
valves would have been an excellent improvement over AMR head
performance. It is difficult to say whether tunnel valves will become a viable
replacement for spin valves. Although virtually all of the recent advancements
in areal density have been associated with large increases in read head
amplitude, long term progress may require radical changes in the media, such
as perpendicular recording schemes or patterned media (Thompson and Best,
2000) .
Throughout the history of magnetic data storage, the end has been
predicted just over the horizon (Eschenfelder, 1970). In fact, however, the
horizon of development is always only a few years. Recently, the technology
seems to exceed demand, and price has become a stronger factor in
differentiating the competition (Levitas, 1999) . However, new video
applications on the horizon will require even more storage than is now
available. It appears that data storage will remain a highly competitive and
rapidly developing industry in the foreseeable future.

Appendix 8. 1 Glossary of Commonly


Used Industry Terminology

ABS - air bearing surface; a lithographically patterned surface topography that


enables the head to hover over the spinning disk at a constant height on a
slim cushion of a few air molecules, which can be thinner than 10 nm.
Magnetoresistive Recording Heads 285

Amplitude- signal from the head (mV).


AMR - Ampl itude - signal-anisotropic magnetoresistance effect; the difference
in electrical resistivity with current flowing parallel to the magnetization
from that with current flowing perpendicular to the magnetization,
expressed as a fraction of the total resistance.
A-parameter - width of a transition written into the media.
AP-pinned-anti-parallel pinned; trilayer of a very thin metallic layer such as
Ru sandwiched between two ferromagnetic layers that are strongly
exchange coupled in an antiparallel orientation, with one layer pinned to an
adjacent antiferromagnet via exchange coupling.
Areal density (Gb/in2 ) - number of bits that can be stored per unit area of a
disk surface.
Asymmetry - difference between amplitudes of pulses with positive and
negative polarities, normalized by amplitude; Asymmetry = (Amp+ -
Amp- )/(Amp+ + Amp-).
Bias point- position along transfer curve with zero externally applied magnetic
field, dictated by the orientation of the free layer with respect to the pinned
layer.
Bias Voltage - electrical potential difference across the head (including
leads); (bias current) x resistance.
BLPN-baseline popping noise; see Fig. 8.6b.
Channel- programmed electronic circuit that converts the head's readback
signal into useful information that can be interpreted by a microprocessor;
also encodes data for writing and interprets servo information.
Cylinder-tracks at a single radius on multiple disk surfaces.
Demagnetization field - equivalent magnetic field that expresses the tendency
of magnetic moments to align antiparallel to one another.
ECC - error correction code; data is encoded with a small amount of
redundancy so that errors can be detected and corrected.
Exchange coupling - strong, short-range, oscillatory interaction between
atomic magnetic moments, also known as the RKKY interaction.
Excitation - fraction of the full transfer curve that is utilized by the applied
magnetic field coming from the media.
Ferromagnetic coupling field - exchange coupling between the pinned layer
and the free layer.
FH-fly height; physical distance between head and disk surfaces.
Free layer - layer in a spin valve which is free to rotate in an externally
applied magnetic field.
Gap - two nonmagnetic insulating layers that separate the sensor from the
shields.
GMR - giant magnetoresistance; large change in resistance in an applied
magnetic field, resulting from spin dependent scattering at interfaces
286 Christopher D. Keener

between magnetic and nonmagnetic metallic materials.


Hard bias - a hard magnetic material deposited at the ends of the sensor to
exert a magnetostatic field, thereby pinning its magnetic orientation
longitudinally.
HDD- hard disk drive; rigid disk drive; also known as a DASD, or direct
access storage device.
HGA - head gimbal assembly; suspension with slider attached.
HSA - head stack assembly; a group of HGA' s swaged together on a rotary
actuator, with a preamplifier mounted at the base.
ISI- Inter-symbol interference; the subtractive effect of magnetic fields from
adjacent transitions, leading to overlap of pulses and a decrease in
amplitude at high linear density. If severe enough, lSI may reach the limit
of the channel's capability to distinguish one transition from the next.
Junction - point of contact between the hard bias and the sensor.
Linear density (kbpi) - number of bits per inch along the circumferential
direction of the media.
Longitudinal- the direction parallel to the current flow and the ASS.
Magnetic instability - Sarkhausen noise or thermally activated fluctuations of
the magnetic moment in a portion of the sensor; resistance, amplitude, or
asymmetry of the head changes over time.
Magnetic spacing - total distance between the edge of the free layer and the
magnetic media on the disk; sum of fly height, carbon overcoats on the
head and disk, and any recession of the sensor away from the ASS.
Mrt- Remanent magnetization multiplied by thickness, usually referring to the
media.
MRW - magnetic read width; the term may be used either for the physical
length of the sensor, or for the width of the magnetically sensitive portion of
the reader. Often the two meanings are distinguished by referring to
magnetic MRW (in spite of the term's redundancy) or physical MRW.
Nanolayer - very thin high-moment layer appl ied to the interface between the
NiFe free layer and the spacer layer, in order to enhance the GMR effect.
OW-overwrite; ability of the write head to write over existing data with new
data. This term usually refers to a specific test in which a high frequency
repeating data pattern is written over a lower frequency pattern, and the
remaining low-frequency pattern (expressed in dS compared with the
original low frequency amplitude) is detected using a notch filter.
Permalloy - Niso Fe20 ; used for its soft magnetic properties, low
magnetostriction, and chemical stability.
PES - position error signal; signal that determines the position of the head
relative to the track center, used as feedback in the servo loop to keep the
head on track.
Pinned layer - ferromagnetic layer that is exchange coupled, or pinned, to an
Magnetoresistive Recording Heads 287

antiferromagnetic layer that provides it with unidirectional anisotropy.


PRML - partial response maximum likelihood; channel algorithm in which the
most probable combination of a series of bits is determined according to the
difference of sampled data from ideal target values that would be observed
in the absence of noise.
PW50 - pulse width at 50 % of the ampl itude; may be expressed in time units
(ns) or distance (nm or IJm), but distance units are more generally
applicable because this expression of PW50 is less dependent on test
conditions.
Random telegraph noise - thermally activated switching between two
metastable states of the magnetization in a portion of the free layer,
resulting in the resistance jumping between two levels at random times.
Resolution - ratio of high frequency amplitude to low frequency amplitude;
normally varies inversely with PW50, and an indicator of the ability to
distinguish adjacent transitions.
RKKY - Ruderman Kittel, Kasuya, and Yoshida; strong, short-range,
oscillatory interaction between atomic magnetic moments, also known as
exchange coupling.
RRO - repeatable runout; periodic variation of the offtrack position of the
head, often related to asymmetric bearings in the spindle.
SAL - soft adjacent layer; soft layer in an AMR head that is saturated by the
current-induced magnetic field and provides biasing for the AMR layer.
Sendust - FeAISi alloy often used for shield material in magnetic recording
heads (see C.H. Bajorek et aI., u.S. Patent 1918554 (1990)).
Servo - feedback system that uses information stored at regular angular
positions on the disk (thus at regularly timed intervals) to allow the head to
seek the correct track and to remain on track.
SER - soft error rate; raw bit error rate, or number of errors divided by the
number of bits written, before error correction is applied.
SH-stripe height; distance from ABS to back of sensor.
Shields - thick ferromagnetic layers that are sputtered or electroplated both
below and above the sensor (separated from the sensor by an insulating
gap) to absorb flux from nearby transitions on the media, so that one
transition at a time dominates the magnetic field sensed by the head.
Slider - a portion of the wafer substrate together with the deposited sensor,
shields, write head, and leads, and a patterned air bearing surface.
SNR - signal to noise ratio; ratio of ampl itude to r/ s of total integrated noise
power measured with a high frequency pattern written to the disk; usually
expressed in dB.
Spacer layer - nonmagnetic conductor layer between the pinned layer and free
layer of a spin valve.
Spin stand - test equipment that includes a rotating hub for mounting a disk or a
288 Christopher D. Keener

stack of disks, a simulated actuator arm for mounting an HGA or HSA and
allowing the head(s) to fly on the disk(s), electronics, a computer, and
software for testing parametrics such as TAA, PW50, and OW, as well as
for more sophisticated measurements such as error rate testing (often
referred to as Guzik tester in reference to the dominant supplier in this
special ized market) .
Spin valve - a GMR multilayer device containing a ferromagnetic layer that is
pinned by an antiferromagnet, a conducting spacer layer, and another
ferromagnetic layer that freely rotates in the applied magnetic field.
Synthetic antiferromagnet - two ferromagnetic layers at close proximity that
are exchange coupled in opposite orientations; known as an AP-pinned
layer when functioning as the pinned layer of a spin valve.
TAA-track average amplitude; peak-to-peak signal amplitude, measured on
a precision test stand or in an HOD as the average amplitude over a
complete revolution of the disk.
Telegraph noise - short for random telegraph noise.
Thermal stability - capability of the head to maintain a nearly constant
amplitude over a long period of time when operating at high temperatures.
TMR - track misregistration; variations of the off-track position of the head due
to noise in the servo system as well as repeatable runout.
Tracks - concentric circles upon which data is written to a disk in an HOD.
Track density (ktpj) - number of tracks per unit of radius in an HOD.
Transfer curve - resistan of a spin valve measured as a function of the
externally applied magnetic field.
Transition- a magnetization reversal in the media.
Transverse- the direction perpendicular to the MRW, into the sensor from the
ABS.
VCM-voice coil motor; the HOD component that drives the heads rotationally
to the correct track position.

References
Arnoldussen, T. C. and L. L. Nunnelley. Noise in Digital Magnetic Recording.
World Scientific Publishing Co. , Pte. Ltd., Singapore (1992)
Ashar, Kanu G. Magnetic Disk Drive Technology. IEEE Press, N. Y. p. 3
( 1997) This book provides an easy-to-read overview of all aspects of
magnetic recording
Baibich, M., J. Broto, A. Fert, F. Nguyen Van Dau and F. Petrof. Giant
magnetoresistance of (OODFe/(OODCr magnetic superlattices. Phys. Rev.
Lett. 61: 2472 (1988)
Beach, R. S. , M. Pinarbasi and M. J. Carey. AP-Pinned Spin Valve GMR and
Magnetization. J. Appl. Phys. 87: 5723 (2000)
Magnetoresistive Recording Heads 289

Bertram, H. Neal, In-depth theoretical approach to recording physics. In:


Theory of Magnetic Recording. Cambridge University Press, London (1994)
Campbell, I. A. , A. Fert and O. Jaoul, J. Phys. C: Metal Phys. Suppl. 1:
S101 (1970); Jaoul 0., I. A. Campbell and A. Fert. J. Magn. Magn.
Mater. 5: 23 (1977)
Campbell, I. A. and A. Fert In: E. P. Wohlfarth, ed. Ferromagnetic
Materials: A Handbook on the Properties of Magnetically Ordered
Substances. North Holland Publishing, N. Y. ,p. 748 - 805 (1980)
Chhabra, D. S., S. A. Bolasna, L. K. Dorius, L. S. Samuelson. Air bearing
design considerations for constant fly height applications. IEEE Trans.
Magn. 30: 417 (1994)
Dieny, B., V. S. Sperious and S. Metin. Magnetotransport properties of
magnetically soft spin-valve structure. J. Appl. Phys. 69: 4774 (1991)
Diency, B. , V. S. Sperious, S. S. P. Parkin, B. A. Gurney, D. R. Wilhoit and
D. Mauri. Giant magnetoresistance in soft ferromagnetic multilayers. Phys.
Rev. B 43: 1297 (1991)
Egelhoff, W.F. P.J.Chen, C.J.Powell, M.D.Stiles, R.D.McMickel, J.H.
Judy, K. Takano, A. E. Berkowitz and J. M. Daughton. Specular electron
scattering in giant magnetoresistance spin valves. IEEE, Trans. Magn. 33:
3580 (1997)
Egelhoff, W. F. et al. Magnetoresistance values exceeding 21 % in symmetric
spin valves. J. Appl. Phys. 78: 273 (1995)
Eschenfelder, A. H. Prom ise of magneto-optic storage systems compared to
conventional magnetic technology. J. Appl. Phys. 42: 1372 (1970), for
example, which predicted an ultimate limit of 2 Mb/in2
Fontana, R. E. Process Complexity of Magnetoresistive Sensors: A Review.
IEEE Trans. Magn. 31: 2579 (1995)
Gangopadhyay, S. Current and future of magnetic heads. Presentation at
Diskcon 2000 Technical Conference, Sept. 19-21. San Jose, CA(2000)
Granstrom, E. and N. Tabat. Limitations of the adiabatic model for ESD failure
in GMR structures. In: Electrical Overstress/Electrostatic Discharge
Symposium Proceedings. ESD Assoc. , Rome, N. Y. p. 180(2000)
Hardner, H. T., M. B. Weissman and S. S. P. Parkin. History dependent
domain structures in giant magnetoresistive multilayers. Appl. Phys. Lett.
67: 1938 (1995)
Hayakawa, Y. Thin-film magnetic head and its production. Japan Patent, No.
JP020924A2 (2000)
Heim, D. E., R. Fontana, H. Tsang, V. Sperious, Y. Gurney, M. Williams.
Design and operation of spin valve sensors. IEEE Trans. Magn. 30: 316
( 1994). Essential introduction to spin valves
Heim, D. and S. S. P. Parkin. Magnetoresistive spin valve sensor with
improved pinned ferromagnetic layer and magnetic recording system using
290 Christopher D. Keener

the sensor. U. S. Patent, No. 5,465,185 (1995)


Hsiao, R. Fabrication of magnetic recording heads and dry etching of head
materials. IBM J. Res. Dev. 43(1-2): 89 (1999)
IBM Travelstar 6GT announcement, Mar. 4 (1998). IBM announces a record-
breaking notebook PC hard drive: first portable storage device shipping to
incorporate GMR technology
IBM Travelstar 30GN announcement, Mar. 27 (200 1). IBM introduces world's
quietest line of high performance 2.5" hard disk drives(200 1) available, at
http://www.storage.ibm.com/hardsoft/diskdrdl/hddnews.htm
IBM Ultrastar 73LZX announcement, Oct. 24 (2000). IBM offers significant
improvements in storage technology: widest range of capacities, richest set
of interfaces, improved performance, and lower power consumption in a
low-profile drive (2000), available at http://www.storage.ibm.com/
hardsoft/diskdrdl/hddnews. htm
Jones, B. A. Theory of exchange coupling in magnetic multilayers. IBM J.
Res. Dev. 42: 25 (1998)
Kanai et al. Spin valve REad heads with NiFe/Cogo FetO Layers for 5 Gbit/in2
density recording. IEEE Transactions on Magnetics 32: 3368 ( 1996)
Kasuya, T. Prog. Theor. Phys. 16: 45,58 (1956). See review in B. A.
Jones, 1998
Keener C. D., M. B. Weissman. Individual domain wall motion in
NiO. 77MnO. 23 observed via resistance fluctuations. Phys. Rev. B 51:
11,463 (1995)
Kirschenbaum L. S. , C. T. Rogers, S. E. Russek, S. C. Sanders. Telegraph
noise in silver-permalloy giant magnetoresistance test structures. IEEE
Trans. Magn. 31: 3943 (1995)
Koui, K. H. Fukuzawa, H. Tomita, H. Fake, Y. Kamichguchi, H. Twasaki, M.
Sahashi. Specular spin valve GMR head using NOL in pinned layer.
Presentation EB-13 at the 8th Joint MMM-Intermag Conference, San
Antonio, TX, (2001)
Kroumbi, M. Magnetoresistive read transducer having hard magnetic bias.
U.S. Patent, No. 5,018,037 (1991)
Langberg, M. Total TV Control. San Jose Mercury News, Mar. 11, p. 1F
(2001)
Levitas, D. The disk drive market: a period of pain. IDEMA Insight.
September/October,p. 7(1999). See IDEMA Insight for regular updates and
a business focus; free subscriptions available from Idema, and articles are
posted on their web site
Mcinerney, B. Surveying micro-positioning technology for advanced disk
drives. Data Storage. August, p. 24 (2000). See Data Storage Magazine
(Pennwell), for regular updates and reviews of technical specialties; free
subscriptions available from Penwell, and articles posted on their web site
Magnetoresistive Recording Heads 291

Monroe, J. The past may be prologue or history may be bunk: reflections on a


perilous quarter. Idema Insight. July/August, p. 6 (2000)
Murdock, E. S. Measured noise in thin film media. In: Arnoldussen T. C. and
L. L. Nunnelley eds. Noise in Digital Magnetic Recording Systems. World
Scientific Publishing Co. , Pte. Ltd., Singapore chapter 3, p. 65(1992)
Nesbet, R. K. Theory of spin-dependent conductivity in GMR Materials. IBM J.
Res. and Devel. 42(1): p. 53(1998)
Pinarbasi, M. S. Stein, H. Gill, M. Parker, B. Gurney, M. Carey, C. Tsang.
Antiparallel pinned NiO spin valve sensor for GMR head application. J.
Appl. Phys. 87: 5714 (2000)
Potter, R.1. Digital magnetic recording theory. IEEE Trans. Magn. MAG-I0:
502 (1974)
Ruderman, M. A. and C. Kittel. Phys. Rev. 96: 99 (1954). See reviem in
B. A. Jones, 1998
Taratorin, Alexander. Characterization of Magnetic Recording Systems: A
Practical Approach. Guzik Technical Enterprises, San Jose, CA (1996)
Thompson, D. A. and J. A. Best. The future of magnetic data storage
technology. IBM J. Res. Dev. 44: 311 (2000)
Tsang, C. H. R. E. Fontana, T. Lin, D. E. Heim, B. A. Gurney, M. L. Williams.
Design, fabrication, and performance of spin-valve read heads for magnetic
recording applications. IBM J. Res Devel. 42: 103 (1998)
Wang, S. X. and A. M. Taratorin. Magnetic Information Storage Technology.
Academic Press, San Diego (1999)
White, R. L. Giant magnetoresistance: A primer. IEEE Trans. Magn. 28: 2482
(1992)
Yoshida, K., Phys. Rev. 106: 893 (1957).See review in B.A.Jones, 1998
Zhang, J. N. Zhu, Y. Huai, A. Prabhaker, P. Pana, D. Seagle, M. Lederman.
Analysis of random telegraph noise in spin valve heads with ultra-thin free
layers. IEEE Trans. Magn. 37: 1679 (2001)
Zhu, J. -G. , Y. Zheng and S. Liao. Patterned exchange stabilized spin valve
heads at very narrow track widths. IEEE Trans. Magn. 37: 1723 (2001)

The author wishes to thank Zhong-Heng Lin, Ken Mackay, Lew Nunnelley, Joseph Smyth,
Ciaran Fox, Daniele Mauri, Jinshan Li, Peter Melz, Sam Luo, Chin-Yu Yeh, Harry Gill, Bob
Beach, David Heim, Mike Salo, Neil Smith, Takao Matsui, Toyomi Ohsawa, and many other
colleagues here at Hitachi for insightful discussions. Thanks to Ed Grochowski for contributing
Fig.B.1 for this work.
9 Magnetic Random Access Memories for
Computer Data Storage
(Properties and Applications of Advanced Magnetic Materials, Advanced Magnetic Materials)

Frank Z. Wang

9. 1 Introduction

A variety of magnetic random access memory (MRAM) technologies have


been explored over a period of many decades. Their advantages include an
unlimited number of read-write cycles, random access to any address,
radiation-hardness, and non-volatility (namely, the state of the memory does
not require periodic refreshing and is maintained even when power is removed
from the memory).
Originally, MRAM had been in use for 30 - 40 years in the form of
magnetic core memory which was the only affordable random access memory,
although very labor-intensive and high-cost per bit, before the advent of static
RAM (SRAM) and dynamic RAM CORAM) in the form of integrated circuits. It
should be noted that while core memory has disappeared from most
computers, there are still a number of applications for core memory in
specialized systems, primarily military and space systems. Early interest
centered on magnetic bubble technology was not successful in either system,
involving an upper layer of permalloy patterned into a T-I bar structure and a
lower layer of a garnet material with a single anisotropy direction
perpendicular to the surface. Magnetic thin films were considered as an
alternative to the core memory. For the past 10 years or so, there has been a
renewed interest in MRAMs by replacing thin film with anisotropic
magnetoresistive (MR) bit structures. However, these non-volatile memory
technologies show comparatively poor performance, and are thus limited in
their applications. The small number of suppliers would be a concern for
systems requiring long-term maintenance or likely upgrade (Jorgensen, 1979;
Daughton, 1992).
The demand for non-volatile radiation-hard memory has continued with
additional pressures for higher density, lower power, higher speeds, and
lower cost per bit. The present efforts to develop mainly three different
versions of non-volatile solid state memories are a response to this. First, an
Magnetic Random Access Memories for Computer Data Storage 293

MRAM using the spin-valve (SV) effect has been shown to have fast switching
signal. Second, an erasable programmable read only memory (EPROM)
employing the Hall effect to detect the fringe field of a ferromagnetic storage
element has also been proposed. The third development under way is so-
called spin-dependent tunneling (SOT) MRAM. The density-independent signal
level means that an ultra-dense SOT-MRAM can be fabricated without any
signal degradation due to its current-perpendicular-to-plane (CPP) structure.
This article will review recent progress in MRAMs and outline the prospects for
future developments.

9.2 MRAM Using Spin-Valve Effect

MRAM using SV effect is evolved from that using MR effect. A spin-valve is


composed of two ferromagnetic (FM) layers (such as Permalloy or Co) with a
spacer layer of a nonmagnetic conductor (typically Cu). In principle in the
sandwich structures exhibiting the SV effect, the resistance is lower when
alternate magnetizations are parallel than when they are antiparallel. The SV
effect can improve MRAM significantly because of its magnetoresistance
change ratio, 4 % - 26 %, in general much larger than the 2 % of Permalloy.
Twice the signal decreases the read access time necessary to resolve this
signal by a factor of four, as illustrated in Fig. 9. 1.

MR

I
I
Threshold value
I
I

Read access time for spin-valve


Read access time for MR
I

o Read access time


Figure 9. 1 Comparison of read access time between SV and MR. Twice the signal
decreases the read access time necessary to resolve this signal by a factor of four.

Figure 9.2 is an MRAM architecture using FM(hard)-conductor-FM(soft)


SV sandwich proposed by Wang and Nakamura (1995a). This memory
requires the SV bit lines and the word lines. In SV sandwich there are two
kinds of ferromagnetic layers which possess different coercivities. As shown in
Fig. 9.3, this memory operates on the general principle of storing a binary
294 Frank Z. Wang

datum in the magnetically-hard layer (Co) and sensing its remnant state by
switching the magnetically-soft layer (NiFe) in such a way that the magnetic
state of the hard layer remains unaltered. The parallel and anti parallel states
between magnetizations will give a different readout.
Spin-valve sandwich

Word current

Figure 9. 2 A SV-MRAM architecture using SV sandwich. The word lines and SV bit
lines. which are orthogonal to each other. are used for selection purpose. The shorting
bars between elements to reduce the bit line resistance are not shown in this figure.

. <=AppIJed small he~


Magnetization

Soft
(NiFe)

Nonmagnetic II
conductor ,,,::......-:- _
(Cu)
Hard
(Co)

State of "0" State of "1"


(a) (b)

Figure 9.3 Schematic storage and readout mechanism of SV-MRAM.

Addressing any desired element in a two-dimensional (2-D) array is a


basic requirement for MRAMs. This is realized by means of the "astroid-
shaped" switching characteristics of a uniaxial anisotropic ferromagnetic thin
film. As depicted in Fig. 9.4, the switching field in the longitudinal direction is
lowered when a transverse field is applied. The word lines and bit lines, which
are orthogonal to each other, are used for selection purposes (Fig. 9. 2). A
word current passes through one of the word lines and a bit current through one
of the bit lines. The bit current is assumed to flow mainly along the
intermediate layer (e. g. , Cu) of SV sandwich due to its large conductivity.
Magnetic Random Access Memories for Computer Data Storage 295

These two currents combine at one and only one intersection point, i. e. , the
addressed element. As depicted in Fig. 9. 4a, neither the word field (current)
with selected value nor the bit field (current) with selected value by
themselves are able to switch the element. In Fig. 9. 4b if both fields act
together, they are able to produce a combined field vector greater than the
switching threshold of the addressed element. In such a way, a writing or re-
writing operation is realized (Wang and Nakamura, 1996a).

Word field Word field

Un-addressed

Bit field Bit field

(a) (b)

Figure 9.4 How to address a desired element in a 2-D array with the aid of the astroid-
shaped switching characteristics? (a) Neither the word field (with selected value) nor the
bit field (with selected value) by themselves are able to switch the element; (b) both
fields act together to produce a combined field vector greater than the switching threshold
of the addressed element.

Magnetization configuration of an SV-MRAM element and its expected


writing and reading timing diagram are shown in Fig. 9. 5. During a write
operation, the sign of the word current will define the stored bit value, "0" or
" 1." In the absence of external field (sleep mode), the alternate
magnetizations are anti parallel , due to the inter-layer antiferromagnetic
coupling, giving small self-demagnetizing fields and small external stray fields
on nearby memory elements, both of which could help with element stability.
During a read operation (alive mode), data readout is performed by
monitoring the voltage response of the SV bit line against a positive word
current, whose value is not sufficient to switch the hard layer, but switches the
soft layer if it opposes the soft layer's magnetization. A different voltage
depending on whether a "0" or "1" is stored, should appear across that SV bit
line. In the experimental pulse sequence (Wang and Nakamura, 1995a), the
sense current applied into the bit line is 5 mA and a sense output voltage of
8 mV appear. The read/write energy is quite low. Furthermore, the test
indicates that a stable readout state involving 3x 108 repeated reversals, by
field cycling, of the soft layer's magnetization is achieved, without gradually
decaying and eventually erasing the hard layer's magnetization. Thus this
SV-MRAM is confirmed to have a non-destructive readout (NORD) property.
296 Frank Z. Wang

Write process : Sleep mode: Alive mode


r-, !Ni@Fe!(ReadproceSS)
o II II ,, - - - 0',
I I :Co ----! n
Word current-1 ~!:!~.!J_~~~ L -

lU iCDl
! HighMR

B;, '""'"JU I

:
~
,

:~
, I
I -- I
:~HighMR~
:: :~ ~
Voltage across : : I
bit line ------------i---------- : '
I I :
I :0 : o
l~
I

,:
,
l, ~ L_~J
,
I
~
~
LowMR
Figure 9.5 Magnetization configuration of a SV-MRAM element and its expected writing
and reading timing diagram.

9. 3 EPROM Using Hall Effect

An PROM using the Hall effect to detect the direction of magnetic moment of a
ferromagnetic storage element has been proposed by Timoshkov et al. , as
shown in Fig. 9. 6. The magnetic perpendicular anisotropy film of high
coercivity is used for data storage, and the Hall element for data reading. The
insulating layer between them is electrical isolation of the functional layers.
The data recording process is realized by magnetizing the perpendicular
anisotropy layer. The magnetic flux B z of this layer penetrates through the Hall
element, and the data read process is carried out by the Hall effect in the
semiconductor material. The time of data reading for such a memory is limited
only by the relaxation time of the Hall effect processes and is equal to 10 -12 -
10 -13 s. The equation for the Hall voltage U x is

U x = rH. lyB z ((I/o) (9.1)


ned
where rH is the Hall factor depending on the prevalencing scattering
mechanism of the carriers, 1. 18 - 1. 93; I y is the sens e current; e is unit
electric charge, 1. 6x10- 19 C; nand d are the carrier concentration in the
semiconductor (data reading) layer and its width, respectively; ( I/ a) is a
Magnetic Random Access Memories for Computer Data Storage 297

correction function, which is equal to 0.90 - 0.95 for the ordinary ratio of the
length to the width of the data reading layer I I a = 2 - 3. a is the width of the
data reading layer, I is the length of the data reading layer.

Magnetic perpendicular
anisotropy layer
Insulating separating layer

I
I
I
I
Access lines i
I
I
I
I
/ - -_ _U.::..o"-- I:
x
I

Iy
Figure 9.6 An EPROM using the Hall effect to detect the direction of magnetic moment of
a ferromagnetic perpendicular storage element (Timoshkov et aI., 1996).

9.4 MRAM Using Spin-Dependent Tunneling Effect

The spin-dependent tunneling (SOT) effect (Julliere, 1975; Boeck, 1998) is a


special CPP GMR effect, based on FM-tunnel-barrier-FM spin-polarized
junctions. The tunneling resistance between two ferromagnetic metal layers
that are separated by a thin insulator depends on the relative orientation of the
magnetization of each layer.
Work by Wang et al. on the MRAM using SOT effect began in 1994 (Wang
and Nakamura, 1995b, 1996b), and led initially to the successful fabrication of
a 2 x 2 bit SOT-MRAM, based on Col Ab03/80NiFe tunnel junctions.
Figure 9.7 is its scanning electron microscopy (SEM) picture and schematic
diagram. The chiplet has 8 contact pads which provides the required flexibility
in connecting the test site to the peripheral select!drive circuitry. The spin-
tunneling junction dimensions varied between 2 and 50 IJm. All the spin
tunneling samples Co ( 100 nm) I AI 2 0 3 (3 - 8nm) 180NiFe ( 100 nm) were
prepared by radio frequency (rf) sputtering with argon. The intermediate
AI 2 0 3 was formed by oxidation in the atmosphere. Uniaxial magneto-
crystalline anisotropy in ferromagnetic films, important both for memory
storage and for the way that a bit is selected, was induced by a magnetic field
applied during sputtering. Contact windows were opened in the insulator in
order for the orthogonal top leads and bottom leads to directly access the
298 Frank Z. Wang

junctions. The top leads are visible as brighter regions in the SEM picture
whereas the bottom leads are darker.

(a)

(c)

Figure 9. 7 SEM micrographs of a 2 x 2 bit SDT-MRAM chiplet showing the test site
containing 8 contact pads and 4 circular tunnel junctions. each 10IJm in diameter. the
schematic cross section (note the CPP structure) and a single tunnel junction bit. The top
leads are visible as bright regions in SEM whereas the orthogonal bottom leads are darker
(Wang and Nakamura . 1996b).

As shown in Fig. 9.8, the structure of the SOT-MRAM is relatively simple,


compared with the semiconductor memories. Having an active memory-cell
transistor, flash electrically erasable programmable ROMs (EEPROMs) typify
semiconductor memories. The minimum element area of a tunneling cell is
4. 8471 2 , where 71 is the resolution limit of the Iithography used to fabricate the
memory elements. At the moment, 71 is unlikely to reduce below 200 nm,
unless there is significant progress in photo-lithography technology (Daughton,
1992). Compared with more than 1071 2 for the minimum size of a memory-
element transistor (also shown in Fig. 9.8), the SOT-MRAM can potentially be
at least twice as dense.
The spin tunneling junction was designed with a circular shape to benefit
the formation of single-domain structure, based on the micromagnetic
simulation result (He and Wang, 1999), as shown in Fig. 9. 9. For each MRAM
element there should exist only two possible states of magnetization, e. g. ,
left and right. Clearly this is the case for an elliptic or circular island. In
Magnetic Random Access Memories for Computer Data Storage 299

Spin-tunneling cell
NiFe/Ab03/Co ~~'--_ _.........

Minimum area of one Spin-tunneling element: 4.842 2


(a)

N-type silicon
e
e
Minimum area of one semiconductor element: 102 2
(b)

Figure 9.8 Structure comparison of the SDT-MRAM and the semiconductor memory. In
(b) a field effect transistor (FET) comprises a source and a drain contact to send current
through a semiconductor channel and a gate control to modulate the carrier density in the
channel. Here A is the resolution limit of the lithography used to fabricate the memory
elements.

detail, for a single domain island one can be sure that the island switches
completely in a proper writing process. A multi-domain state found in the
square patterned island would lead to a miscellaneous logic in a one-bit per
island recording system. This is because more than two states are possible
and after writing, the island might not have switched completely. As a
consequence, the magnetization of the island becomes unstable and the result
of the next writing process may be unpredictable.
An isolated memory element was tested. Figure. 9. 10 illustrates the
R (H) response's minor loops operating in the mode in which only the
magnetically soft layer is switched by applying a field between 1.6 kA/m
and the sample is initially saturated by + 3. 2 kA/m. Note that the MR
hysteresis at zero field represents different magnetization combinations,
thereby yielding different resistance values.
300 Frank Z. Wang

Figure 9.9 Demagnetized states of patterned islands with different shapes.


The brightness stands for the left magnetization direction and the darkness for the right
(He and Wang, 1999).

Figure 9. 10 Measured resistance vs applied field R (H) for Col AI 2 0 3 180NiFe


(Wang and Nakamura, 1995b).

The on-chip parallel write/read operations were successfully performed.


Writing the information involves a switching of the soft layer's magnetization,
ideally without affecting the reference (hard) layer's magnetization. (In a
memory device, independent switching of the magnetically soft layer is
achieved by making the other layer either magnetically hard or exchange-
biased by an antiferromagnetic layer, e. g., MnFe.) Reading the information
only requires a measurement of magnetoresistance. An output voltage change
Magnetic Random Access Memories for Computer Data Storage 301

of 96 mV, as shown in Fig. 9. 11, between the binary bit states against a sine-
wave sense voltage of 160 mV, 1MHz has been observed. This is a
reproducible readout state.
Vout2 for bit 22 VOUII for bit 21
"- /
Voltage input Vin2
(200 mV/div) 1'..7~/I\- /:KV 1'-.. /
~

Voltage output
(240 mV/div)

I I
Longitudinal
scale:500 ns/dis :4- ~ ,~
96mV

Figure 9. 11 Reproducible read waveform. An output voltage change of 96 mV (after an
amplifier) between the binary bit states against a sine-wave voltage excitation (the
above) of 160 mV, 1MHz has been observed. The vertical scale is 240 mV/ div and the
horizontal scale is 500 ns/div (Wang and Nakamura, 1996b).

Ultra-dense magnetic solid state MRAMs will probably be limited by the


minimum signal level. As a comparison, we first deduce the signal level in
conventional CIP type MR- or SV-MRAM. Its signal is given by
Signalc,P = f J rnax i\ l::..p, (9.2)
where f is a factor smaller than one which depends on the mode of storing and
sensing information, J rnax is the maximum current per cross-sectional area
allowed by thermal dissipation and l::..p is the change of resistivity. Thus at the
maximum current density, the signal level is proportional to the element size
i\. As the storage density increases, the signal level will decrease. In
contrast, the signal of SOT-MRAM results in
Signalcpp =f J max t l::..p, (9.3)

where t is the total thickness of the spin-tunneling junction. As illustrated in


Fig. 9. 12, at a maximum sense current density, the signal level of SOT-MRAM
is obviously independent of its cross-sectional area i\ 2. This result is so
attractive that the density of SOT-MRAM should not be limited by signal
degradation. It is suggested that the density-independent signal level is the
biggest advantage of CPP type SOT-MRAM, compared with CIP type MR- or
SV- MRAM.
It is pointed out that the SOT-MRAM's signal remains while the current-in-
plane (CIP) MRAM' s signal drops with increasing density. However, the
302 Frank Z. Wang

1
1/4

Rl=Pocir- R2=Po (a~2)2 =4RI


V1=1R\
V:! = (Jj4) R2 =V;
A epp (current-perpendicular-to-plane) memory element
(a)

a/a.,,~

L ~:-P"~_P";~R'
R\=P -E.-=p ..L
Oat
V1=1'R 1
0 t
(a/2)t
V-=(Jj2)'R =(1/2)1':
2

A eIP (current-in-plane) memory element


2 I

(b)

Figure 9. 12 Halving the dimension of an MRAM cell will cause different results on signal
amplitude. In the CPP design of (a), signal remains; in the CIP design of (b), signal
drops by a factor of 2. Note that both (a) and (b) are at a maximum sense current
density.

Johnson noise with SOT-MRAM may become a limiting factor in scaling the
MRAM element to range in submicron order because this component of noise
becomes increasingly important and may even dominate the total noise. What
matters for practical application is signal-to-noise-ratio (SNR). Even the
larger signals available from SV structures do not make SV-MRAM attractive
for mainstream RAM applications. In order to achieve reasonable memory
array densities many SV elements (of number N) have to be electrically
connected in series which means that the actual signal available when reading
one particular element is MR/ N (see Fig. 9.2 for reference). Thus, a severe
practical problem is encountered that ultra-dense CIP SV-MRAM will be limited
by low SNR. By contrast, as illustrated in Fig. 9. 7, the high MR signal from
the individual junction element can be fully utilized in a cross-point architecture
by connecting each element in parallel. In this case, the SNR in an N x N bit
SOT-MRAM is independent of N, which is normally a big number. As a result,
although SNR in both CIP and current-perpendicular-to-plane (CPP) MRAMs
falls with increasing density, the SNR for the CPP SOT-MRAM remains much
larger than for the CIP type MR or SV MRAM.
So far, the experimental results of SOT-MRAM are only for a low-density
case in that only 2 x 2 bit elements are included. For a dense or ultra-dense
Magnetic Random Access Memories for Computer Data Storage 303

dense MRAM, the crosstalk from an addressed element to adjacent elements


may cause mis-selection and creep problems (after many such disturbances
the magnetic state of these bits creeps either to some intermediate state or
complete reverse). Note also that the driving field required to switch the layer
increases as a result of the decreased length of the element due to the
demagnetization effect. To address these issues, a finite element method
(FEM) was used to study the electromagnetic behavior of SDT-MRAM (Wang
et al. , 1997). Three keepered SDT-MRAM designs have been considered, as
shown in Fig.9.13. A low-cost conventional keeper-less MRAM design
(Fig.9.13a) is taken as our reference. To localize the field, an improved
partially-keepered MRAM design (Fig. 9. 13b) differs from the reference
MRAM by the addition of partial keepers, with high permeability, above the
bit lines and underneath the word lines. The third design called the
continuously-keepered MRAM design, is shown in Fig. 9. 13(c). In this design,
additional continuous keepers with O. 1 IJm thickness are formed. It was
hypothesized that the drive field would be confined within the nearly-closed
keeper effectively to enhance the driving field and reduce interference
between adjacent cells. The continuously-keepered MRAM design was found to
reduce the crosstalk by a factor of five and reduce the driving current by a
factor of four, compared with a conventional keeper-less design, which will be
the most favored approach for achieving 109 bit! cm 2 areal density.

Co
Bit line

(a) (b)

(c)
Figure 9. 13 Three keepered SDT-MRAM designs: (a) conventional keeper-less MRAM
design; (b) improved partially-keepered MRAM design; (c) further improved continuously-
keepered MRAM design (Wang et aI., 1997).
304 Frank Z. Wang

9.5 Future Perspective

Read-access time and storage density are the twin keys to computer data
storage devices. Although new electronic devices, processor organizations
and software systems have contributed to enormous advances in computer
technology, they would have been worthless without the faster and denser
memories that were developed with them. The SOT-MRAM outlined here in
detail, the most hopeful candidate for magnetic nonvolatile memory, is likely
to achieve very high densities. From Fig. 9.14, it can be seen that SOT-
MRAM, which is simply-structured, may in some applications potentially
exhibit excellent properties superior to semiconductor memories and/or
magnetic disks. In particular, the density-independent signal level is the
biggest advantage of CPP type SOT-MRAM, compared with the conventional
CIP type MR- or SV-MRAM, whose signal level is inversely proportional to the
square root of the storage density. Furthermore, it has also been
demonstrated that although the Johnson noise makes a large contribution to the
total noise, it does not significantly degrade the performance of the device
since the SNR for the SOT-MRAM remains much larger than for the CIP MR or
SV MRAM even at high storage density.

ns

f!S

.g"
'"~ ms
~

[ Floppy]
disk

Mega Giga Tera Quadrillion


Capacity (bits/unit)

Figure 9.14 The position of SDT-MRAM in various data storage devices.


Magnetic Random Access Memories for Computer Data Storage 305

Another important parameter for SOT-MRAM is the intrinsic RC time


constant of the device. Because of the insulator between electrodes, junctions
also act as a parallel plate capacitor. The reason that the RC time constant is
important for memory application is because of speed limitations in reading
data from a memory element. The RC time constant is independent of the area
of the junction and exponentially dependent on the thickness of the barrier and
can be reduced until the processing difficulties with a thin barrier prevent
further reduction in thickness. For modern MRAM applications, the equivalent
resistance for SOT devices must be at the lower end of the resistance range.
Parkin et al. found that the resistance-area products of magnetic tunnel
junctions can be varied from 109 Q IJm2 to as low as 60 Q IJm2 by varying the
AI thickness and properly oxidizing it (Parkin et al., 1999). Incomplete
oxidation leads to the presence of metallic AI in the barrier which results in a
rapid suppression of the tunnel junction magnetoresistance. Based on these
facts, the RC time constant for the junctions should be of ns order (Sousa et
aI., 1999). Nevertheless such an SOT-MRAM has to be recovered with a
silicon sense amplifier. This means a mixed silicon-tunneling structure: the
x - y decoders/sense amplifiers in an array in silicon with "holes" to be filled
in by tunneling array. Thus, the speed of the device will probably be limited
by silicon technology. (Today, the speed of experimental silicon devices has
improved to 1 ns and will be expected to dip below this in future.) There is no
significant access-time advantage for the tunneling design as compared to the
bipolar devices.
As well as the continuing developments which are always to be expected
in semiconductor memory technology, it is also interesting to note other novel
technologies, such as the spin-dependent tunneling (SOT) effect. In the
immediate future DRAM will continue as the densest semiconductor memory,
but MRAM using the SOT effect looks set to take the lead in the medium and
longer term and is a very important development for applications where the
nonvolatility, higher density, radiation hardness and lower power are
required. Prototypes are out (Wang, 2000), technology is maturing, and the
advances mentioned in this article would speed up the pace of the application
of MRAM since a microstructured junction might serve as a high-capacity and
low-power substitute for the conventional semi-conductor memories.

References
Boeck,J. O. Switching with hot spins. Science 281: 357-359(1998)
Daughton J. M. Magnetoresistive memory technology. Thin Solid Film 162:
(1992)
He, L. and F. Z. Wang. Size and shape effects of patterned polycrystal/ine
islands. Intermag, HC -09, Kyongju, Korea ( 1999)
Jorgensen, F. The Complete Handbook of Magnetic Recording. Blue Ridge
Summit, PA, USA(1979)
306 Frank Z. Wang

Julliere, M. Tunneling between ferromagnetic films. Phys. Lett. A 54: 225-


226(1975)
Parkin, S. S., K. P. Roche, M. G. Samant, P. M. Rice, R. B. Beyers,
Scheueriein, E. J. O'Sullivan, S. L. Brown, Bucchigano J, D. W.
Abraham, Y. Lu, M. Rooks, P. L. Trouilloud, R. A. Wanner and W. J.
Gallagher. Exchange-biased magnetic tunnel junctions and application to
nonvolatile magnetic random access memory. Journal of Applied Physics
85(8): (1999)
Sousa, R. C., P. P. Freitas, V. Chu and J. Conde. Vertical integration of a
spin dependent tunnel junction with an amorphous Si diode for MRAM
applications. Intermag, HA-03, Kyongju, Korea ( 1999)
Timoshkov, Y., V. Khomenok , V. Danko and V. Kurmashev. The memory
element. Patent of Russia Federation No.2, 036,517 ( 1996)
Wang, F. Z. and Y. Nakamura. A new type of memory using GMR effect.
1995 IEICE General Conference Proc. C-502, Fukuoka( 1995a)
Wang, F. Z. and Y. Nakamura. Perpendicular GMR random access memory
using magnetic tunneling effect. Journal of the Magnetics Society of Japan
19(52): 108-111(1995b)
Wang, F. Z. and Y. Nakamura. Design, simulation, and realization of solid
state memory element using the weakly coupled GMR effect. IEEE Trans.
Magn. 32(2): 520 - 526( 1996a)
Wang, F. Z. and Y. Nakamura. Spin tunneling random access memory
(STram). Intermag, EC-ll, Seattle, USA (1996b); IEEE Trans. Magn.
32(5): 4022-4024(1996b)
Wang, F. Z., D. Mapps, L. He, W. W. Clegg, and Y. Nakamura. Feasibil ity
of ultra-dense spin-tunneling random access memory (STram). IEEE Trans.
Magn.33(6): 4498-4512(1997)
Wang, F. Z. Diode-free magnetic random access memory using spin-
dependent tunneling effect. Journal Physics Letters 25(2000)
10 Magnetoresistive Thin Film Materials and
Their Device Applications

Dexin Wang

10. 1 Introduction

The study of magnetoresistive materials has been much intensified after the
discovery of the giant magnetoresistive (GMR) effect in 1988 (Baibich et al. ,
1988), along with the decades of steady progress made in the anisotropic
magnetoresistive (AMR) materials (McGuire and Potter, 1975; and Lee
et al. , 2000). In the past several years there have been mainly three
additional factors that contributed in further worldwide research and
development in this area. The first is the achievement of > 10% magneto
resistance (MR) ratio at room temperature for spin dependent tunneling (SOT)
materials (Moodera et al. , 1995). The second is the rapid development in the
disk drive industry that brought in and quickly phased out AMR materials, and
thus brought about the needs for using the GMR materials and beyond
(Johnson, 1997). The third is the need and the potential brought up by these
materials for practical nonvolatile memory devices, partly because of the
failure of other types of magnetic memories such as bubble memory, which
excites intensive research and development of magnetic random access
memory (MRAM) technology (Daughton, 1992a, 1992b).
GMR read head is a great example that a given technology can take less
than a decade from its discovery to mainstream commercial use. Spin valve
GMR read head has been the enabling technology in achieving ..... 100 % areal
density growth rate in the past several years. Several companies announced
plans in shipping MRAM products in 2004. Nevertheless, beyond the MRAM
and read heads, there are many other magnetic devices that have been
developed or under development using a variety of magnetoresistive materials
(Daughton, 1997). These include magnetic field sensors (Carey, 1994),
magnetic field gradient sensors (gradiometers), magnetic galvanic isolators
(Wang et aI., 2001a), magnetic switches, a variety of magnetic MEMS
devices, spin injection devices, and more (Lenz, 1990). Some of these
magnetic devices have been tailored to specific applications such as linear and
angular position sensing, current detection, magnetic media detection,
nondestructive evaluation, and mine detection. These devices are also being
308 Dexin Wang

used in a range of industrial control and automobiles. A great example is the


anti-lock break system (ABS) where integrated GMR sensors show significant
advantages over inductive variable reactance (VR) sensors (Smith and
Brown, 1997). It appears that the family of these devices is getting bigger and
receiving wider acceptance in many applications especially new areas, going
beyond simply replacing old generation devices such as Hall sensors. The
trend appears to be better materials properties, miniaturization, and higher-
level integration with more sophisticated IC electronics. There are also
continued demands in reducing power, size, voltage, and increasing speed. It
is possible that magnetoresistive devices will replace some low end flux gate
sensors (Choi et al., 1996; Kawahito et al., 1994). As reflected by many
other areas, the two signatures of the new era research and development are
the depth of each field and the cross-link among different fields. This
interbreeding has been and will continue to be fueled by the fast exchange of
information via the Internet, numerous conferences, workshops and
international interdiscipl inary collaborations.
The materials and devices described here are mainly based on the work
done by the many scientists at NVE Corporation, and some by other
researchers in the field. MRAM and read head devices using magnetoresistive
materials will not be covered in this Chapter. Interested readers can find the
relevant information in other chapters of this book.

10.2 Magnetoresistive Materials

Many types of magnetoresistive materials have been studied by the magnetic


community at large, including AMR materials, GMR materials, Hall effect
materials, CMR (colossal magnetoresistive) materials, EMR (extraordinary
magnetoresistive) materials, and TMR (tunneling magnetoresistive)
materials. Furthermore, new phenomena related to the magnetoresistive
materials are continually being discovered. The underlying physics is different
for each type of magnetoresistive materials (Prinz, 1995) but the application
aspects are similar, that is, the change in electrical resistance in response to
a change in the magnetic field is used to construct a device. It is also noted
that any effect that can cause the change in the magnetization will cause
similar effect as the magnetic field, such as mechanical stress (for
magnetostrictive materials) and spin transfer (for spin injection structures).
The three most important parameters associated with the magnetoresistive
effect are the maximum percentage change in resistance, MR (%), the field
required to achieve the MR, Hsat , and the field sensitivity that is proportional
to the MR and inversely proportional to the Hsat Secondary parameters are
hysteresis, linearity, offset, bipolar vs. unipolar, and others. In applications,
Magnetoresistive Thin Film Materials and Their Device Applications 309

all the relevant properties are essential, such as thermal stability (Sato
et aI., 1997a, 1997b, 1998a, 1998b), manufacturability, availability, etc. In
practice, the relative importance of each parameter depends on specific
applications. AMR has the lowest H sat , but a low MR too, resulting in low field
sensitivity. GMR materials generally have a high MR but also a high H sat ,
resulting in only a moderately higher field sensitivity than AMR (Wang et al. ,
1997). However, its higher H sat provides a wider field range for the device to
operate. SDT has a high MR and relatively smaller H sat , resulting in the
highest field sensitivity. SDT can also be made with very high resistance in a
small area, thus making it ideal for miniature devices and with a requirement
of low power consumption (Wang et al. , 1999a, 1999b).
Speed is also an important parameter for certain applications using any
type of magnetoresistive materials. The three basic considerations are the
switching speed of the magnetization, the speed of the electronics, the
resistance-capacitance (RC) and inductance-catacitance (LC) constants
(Wang, 2001 a; Russek et al., 1999). It is preferred to reduce the rise time
with which the switch threshold can be below what is predicted by the Stoner-
Wolfarth model (He et aI., 1995).
There are several resonance frequencies to be concerned with for near GHz
operations concerning the magnetic layers. The ferromagnetic resonance (FMR)
frequency is one which is proportional to the square root of the magnetic anisotropy
energy, which is usually not far from 1 GHz (Korenivski and Van Dover., 1997).
For each type of magnetoresistive materials, there are many subtypes.
For example, GMR has antiparallelly exchanged-coupled multilayers (Parkin,
1991 a), uncoupled multilayers (Bussmann et aI., 1998), discontinuous
multilayers (Jarratt et al. , 1997), granular films, spin valves (Dieny et al. ,
1991), symmetric spin valves (Anthony et al. , 1994), and some others. Only
these important applications will be discussed in this chapter.
A general comparison of these properties for several types of practical
magnetoresistive materials is given in Table 10. 1.

Table 10.1 Comparison of several practical magnetoresistive materials.


MR(%) H..,(Oe) Sensitivity( % IOe) Comment

Low sensitivity
Hall - - 0.01
Not for low field
AMR 2.0 4-20 0.5 Not for high field
Multilayer GMR 10-100 100- 5000 0.1 Not for low field
Granular GMR 4-40 20-8000 0.2 High noise
Spin Valve GMR 5-20 5-40 2.0 Low resistance
Sandwich GMR 5-10 10-40 0.5 Require scissoring current
SDT 30-50 5-20 6.0 Difficult process control
310 Dexin Wang

10.2.1 AMR Materials


AMR is the effect of anisotropic change of electrical resistance in response to
the rotation of the magnetization relative to the electrical current. The
fundamental reason for this effect is the anisotropic scattering cross-section of
conduction electrons depending on whether they travel along or perpendicular
to the magnetizations. In contrast, the Hall device works on the principle of
Lorenz force, the trajectory of a moving electron bends in a magnetic field,
according to the left-hand rule. One can define two values of resistance, R II
when the current path is parallel to the magnetization and R 1- the value when
the current is perpendicular to the magnetization. Then the resistance R
normalized by its perpendicular value R 1- is
(10.1)
where MR = (R II - R1-) / R 1- is the magnetoresistance value and e is the
angle between the current and magnetization (Bertram, 1994). Note that the
MR is often used as the vertical axis label for magnetoresistance, which is not
to be confused with the MR value when saturated at high field.
One of the best AMR materials is Permalloy (Ni B1 Fe19)' Permalloy has
very high intrinsic magnetic permeability because of its very low anisotropy
field around 4 Oe and relatively high magnetization of 1.0 T. Its low anisotropy
field is a combination of near zero magnetocrystalline anisotropy, near zero
magnetostriction, and low induced anisotropy. Its low but well-defined induced
anisotropy field makes it easy to be used in a controllable way. It has an MR
value up to 3. 5 % in a field range of a few Oe. However, a high MR value can
only be achieved for thick films. There is a strong thickness dependence of the
MR value, with the real physical origin is still a matter of investigation (Katada
et al. , 2000). MR value is about 2% for a film of 20 nm thick. The resistivity
of bulk Permalloy is about 20 IJQ cm. For thin films, the surface scattering
effect causes the resistivity to increase when the thickness approaches the
mean free path of conduction electrons. A thinner film is preferred because it
provides a high sheet resistance and resistance for the same number of
squares of materials. The best results achieved for films of 20 nm thick have
been 4.0 % MR using a NiFeCr buffer layer (Lee et al. , 2000).
Another useful AMR material is NiFeCo (19 %). It is similar to Permalloy
on all other aspects except that it has a higher anisotropy field of 18 Oe,
compared with 3.5 Oe of Permalloy. NiFeCo is used in some cases in order for
the device to be operable in a wider field range. It often shows smaller
magnetization, dispersion, and therefore, less hysteresis in the hard axis. It is
becoming an important soft magnetic material especially in controlling the
effective anisotropy for devices by balancing the effects of shape, induced,
fringe, and others, as will be discussed in later sections of this chapter.
Magnetoresistive Thin Film Materials and Their Device Applications 311

There is a field induced magnetic anisotropy of the order of a few De up to


150 De. This is widely believed to be caused by atomic paring in the body of
the material. The orientation can be changed by applying the field in another
given direction (other than 180) at sufficiently high temperatures for a
sufficiently long time. It is also noted that this anisotropy field exists even
without intentionally applying any magnetic field during deposition or
annealing. That is because the earth's field is always at action. It actually
takes a smaller field. much smaller than the value of the anisotropy field. to
serve this purpose. However. in practice. a much stronger field is appl ied to
overpower the stray fields from inside the film as well as the surroundings
including the earth's field. and to speed up the process of forming a well-
defined magnetic anisotropy.
AMR resistor bridges are being used in magnetic field sensor products.
The barber pole type of configuration is normally used to achieve a push-pull
action for a Wheatstone resister bridge, with a schematic shown in Fig. 10. 1.
Another type is the serpentine structure with adjacent quadrants perpendicular
to each other. However, due to the two-fold symmetry of the AMR effect,
there are two equivalent output responses from the same device for the same
field. as shown in Fig. 10.2. Therefore, resetting circuits, as shown in Fig.l0.3.
is always needed to set the magnetization to the correct orientation at the
beginning, and restoring the magnetization many times in a minute to minimize
the effect of excursion field. This is not the case for other types of magnetoresistive

":'~GND
Easy axis V
oul
tSensitive axis
Figure 10. 1 An AMR resistor bridge with barber pole bias.
15

10

5
>
5 0 ...... Reset
:5
--Set
i -5
~
; -10

-15

-20 ' -_ _- - L --'- ' - -_ _- ' -

-2 -1 0 2
Applied field (G)
Figure 10.2 The bridge output voltages for an AMR bridge sensor with the two polarities.
312 Dexin Wang

r--:::l:::I:::::l:::-.. Applied field .---:\::::l=f=k.. Applied field

Figure 10.3 The set/resetting circuits for flipping the magnetizations used
in an AMR bridge sensor.

materials, as will be seen later. Another drawback of AMR is its relatively low
MR ratio, which is limited to about 2% for a film in a useful thickness range.
However, its simplicity in film structure and relatively high sensitivity make it
very successful in applications for earth-field and lower-field detections.

10.2.2 Giant Magnetoresistive Materials


It was discovered in 1988 independently by two research groups that artificially
synthesized alternating magnetic and nonmagnetic multilayers exhibit
magnetoresistance (Baibich, 1988). Shortly after that, an oscillating
magnetic coupling respective to the nonmagnetic layer thickness was also
discovered (Parkin, 1991 b). The fact that the value of the MR is so much
larger than AMR, in the range of 10 to 100 % vs. 2 %, was named giant
magnetoresistance (GMR). Magnetic multilayers have been used for
manufacturing magnetic sensor devices and digital switches by NVE since 1994
(Simonds, 1995). However, it was not until the spin valve structure was
demonstrated that the industry began investing heavily in the research and
development of the technology. Followed that spin valve discovery was that
people have realized that GMR works without requiring the oscillatory
exchange coupling. The only exception for mutilayers without requiring an
antiparallel coupling is the decoupled thick/thin or high/low Hc layers (Wang
et aI., 1996). Later symmetric spin valve was evolved from the spin valve
structure and becomes successful, too. A variety of materials systems have
been studied including Co/Cu, Fe/Cr, NiFe/Cu, Co/Ru, etc. Co/Cu shows
the highest GMR values.
The basic spin valve structure is shown in Fig. 10.4. NiFe/Co/Cu/Co has
been used most often for applications with a thin dust of Co at the interface
with Cu. When the electrical current path is perpendicular to the film plan, the
so-called CPP GMR is even stronger than CIP GMR, the current parallel to the
film plan. Nevertheless, multilayer materials rely on antiparallel exchange-
coupling to realize the antiparallel!parallel transition. However, due to the
Magnetoresistive Thin Film Materials and Their Device Applications 313

geometrical limitations of the available patterning techniques, it is difficult to


have high enough vertical resistance. Nevertheless, all of the above-
mentioned types of GMR materials are being investigated for device
applications. In the following, the governing equations for applications are
discussed. References are abundant for readers interested in the fundamental
theories on GMR (Zhang, 1998).

Pinning direction

Current

~&Ifi'ld
Pi nni ng layer
Pinned layer
Spacer (Cu)
Free layer
Easy axis of free layer

Figure 10. 4 The basic spin valve structure with a free magnetic layer and a pinned
magnetic layer sandwiching a Cu spacer layer. The pinned layer is pinned by an
antiferromagnetic pinning layer. The arrows depict the magnetizations, current, and
applied fields, showing the orthogonal configuration of the spin valve device.

At any given angle between the two magnetizations for two magnetic
layers separated by a conducting spacer, the resistance R is

R/R p = 1 - ~ GMR (1 - cose) = 1 - GMR sin 2 (e/2) (10.2)

where GMR = (RAP - Rp)/R p is the giant magnetoresistance value, e is the


angle between the two magnetizations of the two magnetic layers, R p is the
resistance when the two are parallel (e = 0) and RAP is the resistance when
the two are antiparallel (e = 180). It is often self-explanatory whether GMR
means the maximum value or a given value under a specific field. It is noted
that the resistance at any given angle is not directly related to the current
direction in the film plan. However, the GMR value is. When the current flows
perpendicular to the film plane of the interface (the so-called CPP
configuration), the GMR value is the highest. When the current flows parallel
to the film plane or the interface (CIP configuration), the GMR is lower due to
the current shunting effect from the conducting spacer. Currently all the GMR
products use CIP fashion but CPP type of devices are also main interests for
research and development, e. g., VMRAM (Bussmann et aI., 1999) and
VGMR heads (Pohm, 2001), with examples given in later sections of this
chapter.
The linearity response of resistance with respect to field comes about
because that the magnetization has such a relationship with field with a uniaxial
anisotropy H k :
Energy = K' sin2 e+ M' H' cos e, by minimizing the energy one
obtains
314 Dexin Wang

2K* 'sinecose-MsH; sine = 0,


2 K / M: cos e = H.
defined as

Therefore,
H k cos e= H.
Plug this into second part of Eq. (10.2) we have
1
R/R p = 1 - "2GMR (1 - H./H k ). (10.3)

Therefore, the resistance is linearly related to applied field H. in an ideal


case. However, due to effects of shape, magnetostatic field, defects, and
multi-domain configurations, the actual MR vs. field can have different forms
and with hysteresis. Micromagnetic simulation has been useful in guiding the
design of the magnetic devices including MRAM and sensors (Gadbois et al. ,
1998). It is the device designer's task to maximize the use of the desirable
aspects and minimize the effects of the undesirable.
Antiferromagnet (AFM) materials had not been useful until the discovery
of spin valves where AFM is used to pin one magnetic layer as a reference
layer. Therefore, it is straightforward to relate the external magnetic field to
the magnetization of the other so-called free layer, to the resistance of the
sandwich structure. There are a variety of selections of AFM materials such as
FeMn, IrMn, NiO, NiMn, PtMn, CrPtMn, PtPdMn, and so on (Lederman,
1999; Shen and Kief, 1996; Qian et aI., 2001). In selecting a specific AFM
material, it is often desirable to have a high coupling strength, a high blocking
temperature, a low treatment temperature, a low conductivity, a small
thickness required, flat temperature dependence, compatible with a specular
scattering approach (Egelhoff et al., 1996, 1997), low cost, and good
compatibility with other processes in the fabrication.

10.2.3 Tunnel Magnetoresistance


Tunnel magnstoresistance (TMR) is also called SDT. The earliest
experimental observation and theoretical prediction of the MR value to the spin
polarization of such effect was reported (Julliere, 1975). For two decades
after that, TMR values have been around a few percent and mostly at lower
than room temperatures. In 1995, Moodera et al. convincingly demonstrated
that this effect can be made as high as 15 % at room temperature and can be
achieved with a relatively low magnetic field (Moodera et al., 1995). The
TMR has been increased every year since that time, and a 49 % TMR at room
temperature was reported by several research groups in 2000. Besides the
Magnetoresistive Thin Film Materials and Their Device Applications 315

high MR value and low saturation field, the thermal stability has been found to
be adequate for device applications (Wang et aI., 1999a, 1999b; Cardoso et
al. , 2000). Most research and development efforts in using TMR have been
concentrated on MRAM applications (Oaughton, 1997; Gallagher et aI.,
1997; Tehrani et aI., 1999). Motorola and IBM announced plans in shipping
1 MB, and 256 MB products in 2004, respectively. NVE has demonstrated
SOT sensors CTondra et aI., 1998), and prototype isolators (Wang et al. ,
2001a). Several magnetic recording companies have made significant efforts
in applying this material as the next generation read heads after GMR.
A typical SOT structure, as shown in Fig. 10.5, has two magnetic layers
separated by a dielectric barrier layer. One magnetic layer either is pinned or
has higher coercivity than the other to allow the two magnetizations to align
either parallel or antiparallel (Moodera et aI., 1995), thus realizing the lowl
high electrical resistance states for tunneling (Alvarado, 1995). The maximum
relative tunnel resistance change (parallel resistance R as denominator) is
twice of the product of the spin polarizations, P and P' , of the ferromagnetic
electrodes divided by the difference of unity and this product, as described by
Julliere's model (Julliere, 1975).

Figure 10.5 A typical SOT structure has two magnetic layers separated by a dielectric
barrier layer. One layer of metalization is shown in making the electrical contacts to the
two electrodes of the junction.

The relative change in tunnel resistance is also proportional to the cosine


of the angle between the two magnetizations. The nominal resistance of an
SOT junction increases dramatically as the barrier thickness. The direct
tunneling current density J as described by Simmons model is (Simmons,
1963):

J = J o { ( ~ - ~ )exp [ - A(~ - ~) 2
1

J- (~ + ~ )ex p [ - A ( ~ +~ rJ}
1

(10.4)
where ~ is the barrier height,and V applied voltage

Jo = 2:~S2 and A = 4~S(2mee)t


316 Dexin Wang

where s is the barrier thickness, e is electric charge of an electron, h is the


planck constant, me is the mass of an electron. This equation can be used for
both spin-up and spin-down electrons with a different barrier height, thus
leading to a dependence of tunneling current on the alignment of the
magnetizations of the two magnetic layers. Oue to the vertical nature of an
SOT junction, it is convenient to use the resistance-area-product (RAP) as a
figure of merit to describe the resistivity of a particular junction.
The heart of an SOT junction is the insulator barrier. An ideal barrier layer
has a uniform physical thickness, a sharp and smooth interface, homogeneous
composition, is impurity-free, has good thermal, chemical and structural
integrity, and a moderately high energy gap. AI 20 3 is almost universally used
for the barrier material by dozens of research groups. The basic fabrication
process is to deposit a thin AI layer of 6 to 20 A after the bottom magnetic
layer deposition, followed by an oxidation step. The variations for different
research groups are in the oxidation step, which can be plasma oxidation,
natural oxidation, glow discharge, atomic oxygen source, or sputter etching in
oxygen partial pressure. Except for sputter etching in oxygen partial pressure,
all the other techniques seem work well with at least some of the systems.
However, there can only be one preferred oxidation technique for any given
specific system.
The junctions are formed either by using shadow masks or by lithography
patterning (Beech et aI., 1996). The latter could be either a self-aligned
process or a two-step etching process. It is conventional to use junction pairs
with the two-step process. For small junctions, especially for sub-l.Im2 sizes,
self-aligned process is normally used.
SOT junctions have many advantages over other types of
magnetoresistive materials. They can be made to have very high field
sensitivity due to their high JMR ratio and little exchange coupling between the
two magnetic layers. They also have a very high intrinsic resistance, leading
to very low power consumption, and/or very small lateral sizes. Critical to all
tunnel properties is the quality of the insulating barrier and its interfaces with
the two magnetic electrodes. There is a waviness of this interface, as shown
in Fig. 10. 6, which is associated with so-called Neel coupling (Neel, 1962;
Kools, 1996; Wang, 2000a).
An example is given in Fig. 10.7 for a minor MR plot showing only the free
layer switching. The sharp transitions at relatively low fields of about 10 Oe
are ideal for digital isolator/sensor applications.
The junction properties are improved after annealing at 275C for 2 h,
suggesting good thermal stability (Wang et aI., 1999a). Further improved
junctions show a > 20 % MR values after the 350 t anneal for 1 hour. This
high thermal stability proves to be useful in integrating the SOT materials with
IC electronics, which will be discussed later in this chapter.
SOT has been advanced greatly since 1995 and has been developed into
technologies close to being commercialized. There have been demonstrations
Magnetoresistive Thin Film Materials and Their Device Applications 317

Figure 10.6 An HRTEM image of a cross-section of an SDT junction. The waviness and
the nonuniformity of the interfaces are clearly seen.

After anneal at 275 'C for 2 h


(148-3,83827cOI)

1580
MR=21.2%
g RAP=362 kQ'llm2
ll) Area=274 11m2
g 1480 Hpin=1750e
;S
Ul
.;;; Hc=90e
~

.g"u 1380
"
: :l
--,

1280 '-=-------,JL-----,-l,------.:------:-':-0----:!2-=-0---'1:-'::-0
-30 -20 -10 0 I J

Applied field (Oe)

Figure 10.7 A minor MR hysteresis loop shows only the free layer switching.

for SOT MRAM by NVE, Motorola, IBM and others, for read heads in HHO by
Seagate, TOK and NEC, and for low field sensors by NVE. SOT devices have
been fabricated with RAP values ranging from 10 0 IJm2 to lOGO IJm2 and
JMR up to 49 % by Nakashio et al. (2000). These devices have lateral sizes
ranging from sub-lJm2 to multiple mm2 Nevertheless, advanced devices
present new challenges for both the barriers and the magnetics. The biggest
challenge for read head application is to make reliable yet thin barriers with
only -6 A AI in order to achieve a <5 O IJm2 RAP for 100 Gb/in2 recording
318 Dexin Wang

density. The challenge for large MRAM chips is to have extremely uniform
barrier properties as well as magnetics. As for sensor applications, the
challenges are to reduce barrier noise and have better control of magnetics.
In particular, the junction properties are extremely deposition system
dependent, even for the same materials with the same techniques, because of
the extreme sensitivity to subtle changes in deposition parameters. There has
been little effort in exploring new barrier materials other than Ab03' Only
limited work has been done on AIN, MgO, HfO, Ge, etc. Searching for a
barrier other than Ab 0 3 with a lower barrier height will definitely be
worthwhile.
NVE has fabricated SOT devices with RAP values ranging from 1 kO IJm2
to 1 GO IJm2 . These devices have sizes ranging from 5IJm2 to 100,000 IJm2 .
The resulting range in resistance is from> 1 GO to < 10.
For linear sensor applications, a special perpendicular biasing technique
can be used. The basic idea is that one of the two magnetic layers (hard
layer) is pinned with an antiferromagnetic pinning layer while the other
magnetic layer (soft layer) is free to rotate in an applied field. A bias field is
applied perpendicular to the easy axis of the free layer. If this orthogonal
biasing field is slightly higher than the anisotropy field, the soft layer will be
extremely sensitive to external fields along the easy axis but still rotate as a
single magnetic domain. By using this orthogonal biasing technique, NVE has
demonstrated SOT elements with very high field sensitivity (about 3%/Oe),
low hysteresis, and a noise floor at high frequency near the Johnson thermal noise
limit. We expect that the sensitivity can be improved to at least 10%/Oe with more
careful perpendicular biasing. The output from a device biased in this way is
shown in Fig. 10.8.
Orthogonally biased SDT element
14

-15 -10 -5 5 10 15

Figure 10. 8 A linear SDT bridge sensor is achieved by biasing the bridge in the
orthogonal direction using an on-chip coil. The field is swept along the easy axis of the soft
layer, and the hard layer is pinned in the negative direction, parallel to the easy axis. The
dotted line, inserted as a guide to the eye, corresponds to a sensitivity of roughly 3 % IOe.
Magnetoresistive Thin Film Materials and Their Device Applications 319

It is important to use more stable magnetic materials at the barrier


interfaces to reduce any interlayer diffusion while preserving a high MR ratio.
The binding energies of Co, Fe, and Ni with oxygen and nitrogen are all less
than AI. The one with the lowest value should be the best for minimal nitrogen
interdiffusion at the interfaces. There is also a tendency for AI to diffuse into
magnetic materials, with different rates depending on the specific magnetic
alloys. A proper design of the interface with the tunnel barrier is to consider
both stability and JMR ratio. FeCo is preferred for the pinned layer due to its
high magnetization, which results in a high JMR, and proper crystal structure
and texture for the top pinning layer of IrMn.
For the pinning materials, we are using both IrMn and CrPtMn to pin one
magnetic layer and have achieved a blocking temperature above 200 C.
CrPtMn has better chemical stability and higher pinning field but requires a
thicker layer of 30 nm, compared with 10 nm for IrMn. All the magnetic layers
that we have tried are well behaved at 150 C and can stand for 300 b
annealing, if the interface is designed properly and the structure is annealed in
a magnetic field.
The simplest theory on TMR is by Julliere:
t:.R _ 2P , P'z
(10.5)
R - 1 - P 1 P'z
where P 1 and P' zare spin polarizations of the two electrodes that are
ferromagnetic. According to this equation, there is no limit how high the TMR
can go because P can be close to unity. Many people are trying to fabricate
tunnel junctions using half metallic materials such as Huesler materials and
CrOz due to their theoretical values of spin polarization of 100 %. However,
preserving its high polarization at the interface with the insulator barrier has
proven to be difficult. Currently this is a hot topic in the research community
and tremendous progress has been made.

10.2.4 Colossal Magnetoresistance


After the discovery of GMR, another effect was renamed colossal
magnetoresistancel (CMR) because the value of MR is even much greater than
GMR (Berger et al. , 2001). This effect exists in one type of material similar
to the high- T c superconductor, which is the perovoskite material. Around the
phase transition temperature range of this material, a magnetic field can
change the critical temperature by degrees. Therefore, at a constant
temperature, a magnetic field can change the material from an insulator to a
conductor, or vice versa. The ratio of resistance is thousand-fold, and
normally requires a huge magnetic field of many tens of kOe and often at low
temperature. Nevertheless, its share of high MR value interested many
researchers into the field. Currently, room temperature materials and lower
320 Dexin Wang

saturation field materials have not been achieved. However, this type of
material still suffers from the drawbacks of being unstable thermally, difficult
to fabricate (where annealing temperature can be 1200 to 1363 K for tens of
hours for bulk materials), having narrow temperature range, and less
sensitivity. It will take more progress in most of the areas before they become
viable candidates for commercial applications.

10.2.5 Advanced Magnetoresistive Materials


There are several emerging materials/structures exhibiting unique
magnetoresistive properties, but have not yet received wide attention towards
applications. One example is the extraordinary magnetiresistance (EMR),
which is based on the Hall effect but with very a clever use of a conductor with
it (Solin et al. , 2000). A metal strip is placed next to a semiconductor Hall
slab. Under zero fields, the current mostly flows through the metal strip. When
a magnetic field is applied, the Lorenz force pulls the electrons away from its
path in the metal strip into the semiconductor, till all the current flows in the
semiconductor. Thus the effective resistance of the device could change in
orders of magnitudes. The difficulties of this device are the fabrication issues,
especially the requirement for intimate contacts at the metal/semiconductor
interfaces which lie perpendicular to the substrate surface. According to
theoretical calculations, the scaling will not be very good when the thickness
of the conducting stripe is approaching to the electron mean-free-path.
However, the device already available may be quite sufficient for many
applications in terms of signal size, which is much greater than any other
practical magnetoresistive technologies including GMR and TMR.
Synthetic antiferromagnet (SAF) is really a clever use of the strong
antiferromagnetic exchange coupling found in some FM/NM/FM systems. The
one with the broadest use is FM/Ru/FM and the coupling can be as strong as
several kOe (Leal and Kryder, 1998). When combined with a regular
antiferromagnetic layer, SAF is the material of choice providing the strongest
pinning strength and widely used as pinned layers in spin valve GMR materials
and SOT materials. Re has been found to act similar as Ru. It was also
reported that a Co/Cu/Co system was used in an SAF structure but with a less
coupling strength of about 200 to 300 Oe. Another benefit of using an SAF is its
much reduced fringe field due to the free pole compensation of the two FM
layers when the films are patterned into devices. In addition, the thickness of
two FM layers can be adjusted so that the combined effect on the free layer is
minimized. Research extends the concept to include an SAF for both the
pinned layer and free layer to realize desirable magnetic configurations.
One recent discovery in this area is the magnetoresistive switch effect in
metal/semiconductor granular films where a thousand-fold change in resistance
has been observed when a magnetic field is applied (Akinaga et al. , 2000).
Magnetoresistive Thin Film Materials and Their Device Applications 321

However. the practical ity sti II needs to be improved. and so does the theory
to explain the phenomena.
More advanced magnetoresistive materials also include spins in
semiconductors (SPINS). as the next generation of spintronics materials.
There are extensive research activities around academia and national labs. in
searching for room temperature materials and practical and useful devices.

10.3 Magnetoresistive Devices

There is a wide variety of magnetic fields sensing devices that employ


magnetoresistive materials. The basic configuration for such a device usually
consists of a Wheatstone bridge with four legs of magnetoresistors. Depending
on the specific types of the magnetoresistive materials and sensor designs.
either two or all of the four legs of the resistors can be magnetoresistive. For
better matching at different temperatures and easier fabrication. all four
resistors are usually made of the same magnetorestive material. sometimes
with two of them inactive to the external field by shielding them with a thick
soft magnetic material with high permeability such as Permalloy. Nevertheless
all four legs of resistors can be made active with special construction of the
resistors. with examples given later in this chapter. Once a Wheatstone
bridge is constructed. either a constant supply of voltage or current is provided
across to the two supply nodes on the bridge. The other two signal nodes are
linked to an electronic comparator or more complex amplifiers for read out
and/or further signal processing. as depicted in Fig. 10.9. with SDT junctions
pairs as the magnetoresistors.

SOT-A

SOT-B

SOT junction pair

~~=~~;~~~~2TI Barrier
Pinned layer
Free layer

Figure 10.9 Schematic of a typical SDT bridge device. When a field is applied to the
SDT bridge. an output voltage from the output nodes is fed directly into an opamp or other
more complex electronics. Each leg in the bridge has a series of junctions with a pair
depicted at the bottom.
322 Dexin Wang

In zero magnetic field, the four resistors can be designed to have a


balanced resistance value, thus no signal is from the output nodes. Sometimes
an offset is designed based on the requirement of the electronics. When a
uniform magnetic field is applied across the bridge, two of the four resistors
will have a higher resistance than the other two. Therefore, an imbalance from
the two output nodes is generated and fed into an electronic comparator to
bring about a "1" or "0" to the comparators'outputs, thus a digital magnetic
field sensor device is formed. Such a device is depicted in Fig. 10. 10 with
layout schematic on the left, and the bridge action schematic on the right.

External field

Figure 10. 10 Configuration of four magnetoresistors in a Wheatstone bridge. The two


sides of the flux concentrator are shown as the two rectangular boxes. Flux concentrates
are used to enhance the external magnetic field to R 1 and R 4 in the gap, as well as shields
for the other two resistors, namely R z and R 3 It is indicated on the right that R z and R 3
are shielded.

This simple device can be tailored for many applications such as magnetic
field sensors, field gradient sensors, electrical current sensors, galvanic
isolators, and many others. There are many more derivative devices based on
these basic devices. It should be made clear that not all sensor devices need
the shielding of two resistors in the bridge.

10.3.1 Magnetic Field Sensors


Magnetoresistive sensors are most widely used as magnetic field sensors. The
magnetic field is assumed to be uniform across the entire area that covers all
the four magnetoresistors. Under any given field, all the four resistors will
have the same resistance values without special construction that results in no
output from the bridge. Therefore, magnetic shields are used to cover two of
the resistors. The shields can also be used to concentrate the external field to
the other two resistors. For spin valve material, a special pinning direction
setting scheme (Dovek et a!., 1996; Spoong et a!., 1996) was proposed to
achieve the opposite polarity for two of the resistors relative to the other two,
thus all four resistors become active. The same scheme can be used for spin
valve-like SDT structures. Another approach for SDT is a special use of the
so-called synthetic antiferromagnet scheme (Wang et a!. , 2002).
Magnetoresistive Thin Film Materials and Their Device Applications 323

It is noted that both of these schemes, as well as others similar,


inevitably increase the complexity of fabrication. However, the added
advantages such as doubling the signal, without having to use flux
concentrators/shields, lead to advanced devices, that are under active
development.
Multilayer GMR materials are also used to construct sensors in
Wheatstone bridges, as shown in Fig. 10. 10. The two sides of the flux
concentrator are shown as the two rectangular boxes. These two soft magnetic
slabs are used to enhance the external magnetic field to the two resistors R 1
and R 4 in the gap. The slabs are also used as shields for the other two
resistors, namely R 2 and R 3 The effect of flux concentration depends on many
factors, the gap 0 1 , shield length O 2 , thickness of the Permalloy, and its
permeability.
Typical field dependence of the resistance of the materials is shown in
Fig. 10. 11 at several different temperatures, with a constant voltage supply to
the bridge on the left and constant current supply on the right. Because the
multiplayer material is unipolar with respect to the sign of the applied magnetic
field, shield has to be used null two resistors and also used as flux
concentrators. Sensors made of this material have been on the commercial
market for half a decade (Prinz, 1995).
1.2
Constant current supply
-50C
OC
+SOC
+lOOC
;;-
~
:::>
So
:::>
0
""0
'"
.~
t<i
E
0
Z

-0.4 Applied field (Oe)

Figure 100 11 Magnetoresistance of the multiplayer GMR materials at several


different temperatures.

AMR field sensors are constructed similarly. Because of its sinusoidal


square relationship with the angle has a 180-degree period, the push-pull
action of the four legs of resistors can be realized easily. One such device has
four serpentine resistors each occupying a corner on a square die. It is noted
that adjacent resistors have their strip length perpendicular to each other,
causing a phase shift of 90' respective to the field. Therefore, a push-pull
bridge is naturally formed. However, AMR device need a special resetting coil
324 Dexin Wang

to ensure the functioning of the device because that the materials is intrinsically
unstable under field excursions of even a few Oe. Recently, Honeywell
introduced a new magnetic field sensor using AMR materials. The HMC1052 is
a two-axis sensor on one chip for compassing and position sensing applications
requiring small size, low power and high performance. Advantages of this new
design include nearly perfect orthogonal two-axis sensing in a 3 mm x 3 mm x
1 mm 10-pin miniature surface mount package (MSOP). The HMC1052 has a
sensitivity of 1 mV/ (V Gauss), a field range up to 6 Gauss and can
operate on a supply voltage as low as 1. 8 V. Nearly perfect orthogonal two-
axis sensing and matched sensitivity on both axes improves performance and
accuracy of this sensor. However, on-chip set/reset straps represent the
inherent drawback of using AMR materials.

10.3.2 Magnetic Field Gradient Sensors


A magnetic field gradient sensor, also called a gradiometer, senses the
magnetic field gradient either in the field direction or perpendicular to the field
direction. The actual signal is proportional to the difference in field values
between two sets of magnetoresistors. Due to the accuracy requirement of this
spacing, the two sets of magnetoresistors are often designed on the same
chip. The guiding equation for a gradiometer is that the signal is proportional
to the field gradient, the spacing, and the field sensitivity of the two sets of
magnetoresistors. One can also think of two field sensors separated by
distance, and when a differential signal is taken from these two sensors due to
the field variation, a gradiometer is formed. However, a gradiometer has a
much simpler configuration than two field sensors.
A great advantage for a gradiometer is its immunity to a uniform external
field. This is an especially useful feature for NDE and other applications where
the devices operate in the presence of the earth's field.
Although a gradiometer outputs a signal according to a field differential,
the individual resistors are all experiencing the absolute magnitude of the field.
Therefore, the dynamic range for each resistor is still a critical specification
parameter in determining the environment it can sense. Therefore, the
dynamic range for this type of device needs to be greater than the earth's
field, or whatever the environment field is. Gradiomers are very useful in
detecting anomalies in a presence of a uniform magnetic field, either from the
earth's field, or a coil generated pulse field. Gradiometers are being tested in
anti-lock break systems in an advanced model of cars.

10.3.3 Electrical Current Sensors

A magnetic field sensor can also be used to detect the magnetic field created
Magnetoresistive Thin Film Materials and Their Device Applications 325

by a current carrying wire. A magnetic field sensor can be tailored to a variety


of designs specifically for electrical current sensing including intrusive and
nonintrusive types. The nonintrusive design is preferred for the convenience of
applying to other circuits. In this case, a magnetic sensor is measuring the
magnetic field generated by the electrical current, or the Orster field.
Figure 10.12 depicts such a current sensor. With proper calibration and/or
enough knowledge about the geometries of the setup, the absolute value of the
current can be detected without disturbing the actual current to be measured.
Intrusive devices can be made to measure small current to better accuracy
because the current can be brought into close proximity to the sensors.
However, the diversion of current into the sensor makes it more complex for
the current using circuits.

Axis of sensitivity - - i f=rfl.4=fl-f"'il kn Ir-T-----

Direction of current fl ow
Figure 10. 12 An electrical current sensor using a GMR bridge sensor sensing the field
generated from the current, which is properly orientated relative to the current carrying wire.

10.3.4 Integrated Magnetic Sensors


Advanced magnetoresistive devices combine both sensing elements and
service electronics such as amplification and signal conditioning. Voltage
regulating, feed back and others started to make an impact on several
applications (Smith and Brown, 1997). Magnetoresistive devices are made
with sputtering and photolithography techniques, which are compatible with
semiconductor processing. The devices can be fabricated directly on top of the
IC electronics. In fabrication, surface topology of IC wafers has been a big
issue to accommodate back-end magnetoresistive layer deposition. One way
of avoiding the rough topology has been to reserve dedicated areas on the IC
wafers with no transistors underneath. However, this often reduces the device
area efficiency greatly. Another approach has been to planarize the IC wafer
surface by chemical-mechanical-polishing (CMP) or polymerization. Both
approaches are being used in prototype devices. In fact, MRAM devices are
the best examples of integrated MR/IC integration.
326 Dexin Wang

A magnetic digital switch is designed and fabricated in such a way that it


mimics a reed switch. The devices are being used for cylinder position sensing
in industrial control applications.
Typical electronics functions to be integrated with magnetoresistive
devices are amplifiers, electronic hysteresis for controlled switching, voltage
regulators, delay circuits, ESD staging functions, dividers, logic output, and
others. For low cost applications, two wire devices are preferred over three
or four wire devices because of the cost considerations. The advantages for
integrated single chips are mostly cost, as well as reliability and compactness
to a less extent. Cost is especially critical in the automobile industry where
competition is so severe that often a technology will not be adapted unless it is
cost competitive with existing technologies. In many cases, one chip approach
is more reliable, has smaller total chip area, and smaller package size.
Hysteresis is sometimes built in to avoid false triggering due to noise. It is also
better to amplify the low-level signal close to where it is generated for the
least leads interference from the sensor to the amplifier. The signal is
sometimes converted to digital after enough magnitude is achieved to minimize
this type of electronic noise.

10.3.5 Galvanic Isolators


Data transmission has become more vitally useful as the world becomes more
computerized, connected, and automated. It is desirable, and sometimes
necessary to have ground isolation between different stages of data
transmission. This signal isolation is for over-voltage protection, noise
reduction, and avoiding signal saturation. In the past, the galvanic isolation
has been achieved mainly using optical couplers at low frequencies and
transformers at high frequencies.
Spin valve GMR isolators has been introduced to the commercial market
in 1999 by NVE. The device has a magnetic field generating coil on one side of
an insulator (galvanic barrier) and GMR spin valve on the other side (Hermann
et aI., 1997), as shown in Fig. 10. 13. The transmitting medium is a magnetic
field instead of the photon, as in the case of opto-couplers. A schematic is
given in Fig. 10. 14b for such an isolator and Fig. 10. 14a shows its opto-
analogies. The core component is a GMR bridge with a differential output when
a common current is provided. There is also a signal circuitry called a
receiver, sometimes used as well as a driver piece depending on the bridge's
mode of operation. There are two modes of operations for the GMR resistors:
one is a digital mode where symmetric magnetic hysteresis and two different
resistance values at zero coil current are essential; the other is an analog
mode where a single value of resistance at any given coil current is required.
For the digital mode, an incoming logic signal will go through a driver chip
which translates the" 1" to "0" transition into a positive pulse, and "0" to "1"
Magnetoresistive Thin Film Materials and Their Device Applications 327

Passivation
layer

I (current)

Si substrate with GMRsensor


active circuitry

Figure 10. 13 A schematic of a spin valve isolator showing the dielectric barrier
separating the field generating coil and the GMR bridge.

LEDft Signal transmitted


by photon
Optical c--
isolator I in ----->
Galvanic isolatIOn by Photo detector
bulk magnelic field
(a)

Planar Signal transmitted


coil ~bY magnetic field
Magnetic. -.!i-
isolator I In _

Galvahic isolation
by thick dielectric film
(b)

Figure 10.14 The functional diagram of a GMR galvanic signal isolator. The schematics
are similar for an opto-isolator (a) and a spin valve isolator (b).

transition into a negative pulse of very short width, typically 2 - 4 ns. This
driver chip is basically doing duty cycling, thus cutting down the power
consumption on the input side of the isolator. This power saving can be very
significant at low speeds.
For the analog mode, the input signal and output signal are directly
correlated. The coil can be directly connected to the output such as a bus
system, thus eliminating the need for a separate driver chip, and can operate
at a very high speed. The down side is that it takes power to maintain a slow
changing of signals, that is, a current is always needed to maintain a "1"
signal. An added advantage for the analog mode is that it is intrinsically fail-
safe. In other words, under any event of an input power off, the device will
328 Dexin Wang

always be at a known state. Currently, both modes are being developed, with
most of the product shipping using digital mode.
The single channel isolating devices can be combined to multi-channel
receivers and transceivers. The advantages in low power and small size are
more significant for these multi-channel devices. An SOT isolator device has
been fabricated and tested to be functional (Wang et al. , 2001a). An image
of such an SOT isolator device is given in Fig. 10. 15. As can be seen from and
the waveforms shown in Fig. 10. 16, the driver input is the logic signal from the
function generator. The driver outputs a positive or negative pulse, depending
on the rising or falling of the logic input waveform, with a pulse width of about
4. 4 ns. The receiver output replicates the driver input waveform.
Photo of an integrated SDT isolator

SDT
bridge

Driver
input

Receiver
output

Figure 10.15 Photo images of a driver IC chip, an integrated receiver IC chip with an
SDT bridge. These two chips would be packaged in a single leadframe to form an opto-
equivalent part.
Integrated SDT isolator
Tek stop: 500 MS/s 5 A cqs
'--------------f-..----I---f-.-)


l!.: 186 ns
/!woll""-"~~@: -30 ns

Driver input l
4~ F---"""'' ' ' ' '
. MI +-H+-' H+"'-I+1\:; +i+H+t++-:-IH-HHH+,t-!++!+-:H-+I++'H
DrIver output
,-
t;-

Receiver output
3~
l ~---~~-I
I
Ch3 2.00V Ch4 2.00V MIOOnsChi f 1.76V7Dec2000
AdTechilsolator Math] 2.00V lOOns 18:16:53
4/26/01
Figure 10.16 The waveforms of an input signal to a driver IC chip, the input to the field
generating coil, and the output from the receiver IC chip.
Magnetoresistive Thin Film Materials and Their Device Applications 329

The isolation voltage is normally specified at 2500 Vdc. Because of the


high isolation voltage, it is difficult to combine both pieces of IC electronics,
the driver and the receiver, onto the same chip. However, future development
in semiconductor technology, such as deep trench approach and stack chip
bonding technique, may provide new opportunities for further integration.

10.3.6 Advanced Magnetoresistive Devices


The GMR transistor device reported by Monsma et al. was the first of its kind
in using a Co/Cu multilayer as the metal base in an S/M/S device, with the
emitter chip mechanically bonded to the Si!M substrates (Monsma et al. ,
1995 ) . The voltage modulation has been observed but current was
impractically too low at low temperatures. Many more attempts have been
made to achieve room temperature operation, and to improve the current level
as well as current gains. At this date the device is not ready for real world
appl ication.
Other spintronics devices are also proposed. Currently, the main efforts
have been concentrated on ferromagnetic semiconductors devices, primarily
due to their long coherence length of spin packets traveling for hundreds of
microns, and spin packets' lifetime of nanoseconds. However, it is still a
challenge to achieve an FSC with a Curie temperature above room
temperature. Also, the exact device form, attributes, and applications still
need to be worked out.
A vertical GMR cell and circumferential remnant magnetizations have been
proposed for ultra-high density MRAM. It has been observed that CPP GMR is
higher than CIP GMR for multilayer structures. The scaling to smaller cells for
high density favors the CPP far more than CIP, mainly for geometric reasons.
Therefore, CPP structures are being attempted for MRAM applications. The
most promising type so far is the ring type cell consisting thick and thin
alternating ferromagnetic layers with different switching thresholds (Bussmann
et al. , 1998; Zhu and Zheng, 1998), as depicted in Fig. 10.17. With a clever
design of the braded word lines, it is possible to have a 2-0 array selection for
the VMRAM, as shown in Fig. 10. 18. These are based on some earlier works
on rectangular cell designs (Vavra, 1995). However, significant product
development will be needed to offer this device for commercial use. The
ultimate limit may be the curling of the magnetizations in the ring shape that
goes against the exchange interaction. When the diameter is sufficiently small ,
the interatomic Heisenburg exchange energy within each magnetic layer will
overrun the benefit of eliminating free magnetic poles. In theory, the diameter
can be as small as 20 nm. It will depend on the relative progress of the several
promising types of designs, most notably, SOT cells, pseudo-spin valves, and
latch cells.
SOT latch cell design is very similar to the semiconductor static random
330 Dexin Wang

Hard magnetic
layers

Soft magnetic
layers

Figure 10. 17 A circular MRAM cell with alternating soft and hard magnetic layers with
opposing magnetizations as depicted at the bottom.

Figure 10. 18 A circular MRAM cell with the braided word lines above and below the cell
for 2-D selections.

access memory (SRAM) cell as shown in Fig. 10. 19, in which five transistors
are used for each cell (Tehrani et aI., 1999). It is relatively large but can be
very fast. This design will be very useful for high-speed small memory
applications (Wang, 2001 b). In niche applications, a small, low cost, fast,
and nonvolatile memory is needed.
Supply

Low High

Figure 10.19 The schematic of a latch cell using either a spin valve or an SOT half bridge.
Magnetoresistive Thin Film Materials and Their Device Applications 331

10.3.7 Applications of Magnetic Field Sensors


The magnetic field sensors including GMR and AMR are being used in
commercial market for angular or linear position sensing; current sensing;
handheld wireless appliance applications such as mobile phones, personal
digital assistants, walkie talkies, watches and global positioning system
(GPS) receivers; and medical devices. Figure 10.20 gives a pin out diagram
for a low fields SDT sensor. These sensors are amounted in a variety of
packages such as the S08 type shown in Fig. 10.21. These sensors are used
as gear teeth senors, as shown in Fig. 10.22. Figure 10.23 shows a schematic
of a GMR sensor positioned to measure displacement relative to a permanent
dipole magnet. Sensitive axes are indicated and the components of field along
the sensitive axes for two sensors are graphed. Many new applications are
emerging as the sensor technology becomes more developed, as well as
increasing demands in other areas that may benefit from advanced
magnetoresistive devices.

Pin 8 Pin 7 Pin 6 Pin 5

V+ OUT+ oc- oc+

Orthogonal t
NVE field coil t

DD
OUT- y- pc- pc+
Pin 1 Pin2 Pin 3 Pin4

Figure 10.20 The pin out diagram of an SDT low field sensor prototype.

Vehicle disturbance in earth's field applications for vehicle detection can


take several forms. A single axis sensor can detect if a vehicle is present or
not, as depicted in Fig. 10.24. The sensing distance can be 15 m away,
depending on its ferrous content of the vehicle and the sensitivity of the
sensor. This application may be useful in parking garages in directing entering
332 Dexin Wang

Pin-out
V4 (supply) OUT+

Orientation
chamfer
OUT- V-(ground)
Axis of sensitivity

Figure 10.21 The schematic of a packaged GMR bridge sensor.

Magnet
Spacer
Sensor

(a)

(b)

Figure 10. 22 Side view schematic of a gradiometer (8 pin device) with a biasing
magnet. without (a) and with (b) a ferrous object present.

drivers to where the most available spaces to park are, and other space and
metering management ability if combined with other devices. Another use is to
detect traffic flow at designated cross roads or highways for traffic control in a
single check point to metro-traffic management. In traffic control application,
two sensors could be used to detect the presence, travel direction, and speed
of a vehicle.
Mine detection is another emerging area in using magnetoresistive low
field sensors with high sensitivity. Such SDT field sensors have been
successfully field-tested (Wold et ai., 1999).
There are continued demands in reducing power, size, voltage, and
increasing speed in many other applications. It will be possible that
magnetoresistive devices may replace some low end flux gate sensors in the
Magnetoresistive Thin Film Materials and Their Device Applications 333

Figure 10.23 Schematic of a GMR sensor positioned to measure displacement relative to


a permanent dipole magnet. Sensitive axes are indicated and the components of field along
the sensitive axes for the two sensors are graphed.

Figure 10.24 A representation of vehicle disturbance in earth's field applications. A field


sensor can therefore be used to detect if a vehicle is present.

near future. Because of its low cost and it easy to use, magnetoresistive
sensors will open doors for more applications, and displace existing ones.

References
Akinaga, H., M. Mizuguchi, K. Ono and M. Oshima. Room-temperature
thousandfold magnetoresistance change in MnSb granular films:
Magnetoresistive switch effect. J. Appl. Phy. 76: 357 (2000)
Alvarado, S. F. Tunneling potential barrier dependence of electron spin
polarization. Phys. Rev. Lett. 75:513 (1995)
Anthony, T. C. , J. A. Brug and S. Zhang. Magnetoresistance of symmetric
spin valve structures. IEEE Trans. Magn. 30: 3819 ( 1994)
Baibich, M. N. , J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P.
Etienne, G. Parker, A. Hutten and G. Thomas. Phys. Rev. Lett. 68: 3745
334 Dexin Wang

(1988)
Beech, R. S. , J. Anderson, J. M. Daughton, B. Everitt and D. Wang. Spin
dependent tunneling devices fabricated using photolithography. IEEE Trans.
Magn. 32: 4713 (1996)
Berger, A., J. F. Mitchell, D. J. Miller, and S. D. Bader. Evidence for
competing order parameters in the paramagnetic phase of layered
manganites. J. Appl. Phy. 89: 6851(2001)
Bertram, H. N. Theory of Magnetic Recording. Cambridge University Press,
(1994)
Bussmann, K., S.F. Cheng, G.A. Prinz, Y. Hu, R. Gutmann, D. Wang, R.
Beech and J. Zhu. IEEE Trans. Magn. MAG-34: 924(1998)
Bussmann, K., G. A. Prinz, S. F. Cheng, J. Zhu, Y. Zheng, J. M.
Daughton, R. Beech, D. Wang and R. Womack. InterMag. 1999, digest
GA-03 (1999)
Carey, J. Magnetic field of dreams. Business Week 4: 118 (1994)
Cardoso, S. , P. P. Freitas, C. De Jesus and V. Soares. High thermal stability
tunnel junctions. J. Appl. Phy. 87: 6058 (2000)
Chen, L. H., W. Zhu, T. H. Tiefel, S. Jin, R. B. van Dover and V.
Korenivski. Fe-Cr-Hf-N and Fe-Cr-Ta-N soft magnetic thin films. IEEE
Trans. Magn. 33: 3811 (1997)
Choi, S. 0., S. Kawahito, Y. Matsumoto, M. Ishida, Y. Tadokoro. An
integrated micro fluxgate magnetic sensor. Sens. Actuators A. A 55: 121-
127(1996)
Daughton, J. M. Thin Solid Films. 216: 162(1992a)
Daughton, J. M. Giant magnetoresistance in narrow stripes. IEEE Trans.
Magn. Vol. 28: (1992b)
Daughton, J. M. Magnetic tunneling applied to memory. J. Appl. Phy.
81(8): 3758-3763 (4) (1997)
Dieny, B., V.S. Speriosu, S.S.P. Parkin, B.A. Gurney, D.R. Wilhoit and
D. Mauri. Phys. Rev. B 43: 1297(1991)
Dovek, M. M., R. E. Fontana, V. S. Speriosu, and J. K. Spoong. Bridge
circuit magnetic field sensor having spin valve magnetoresistive elements
formed on common substrate. US Patent, 5,561,368 ( 1996)
Egelhoff, W. F., P. J. Chen, C. J. Powell, M. D. Stiles and R. D.
McMichael. J. Appl. Phys. 79: 8603 (1996)
Egelhoff, W. F., P. J. Chen, C. J. Powell, M. D. Stiles and R. D.
McMichael. IEEE Trans. Magn. MAG-33: 3580 (1997)
Everitt, B. A., A. V. Pohm and J. M. Daughton. J. Appl. Phys.81: 4020
( 1997)
Gadbois, J., J. Zhu, W. Vavra and A. Hurst. IEEE Trans. Magn. MAG-34:
1066(1998)
Gallagher, W. J. ,S. S. P. Parkin, Y. Lu, X. P. Bian,A. Marley, K. P. Roche, R.
A.Altman,S.A.Rishton,C.Jahnes,T.M.Shaw and G.Xiao. J. Appl. Phys.
81: 3741 (1997)
Magnetoresistive Thin Film Materials and Their Device Applications 335

He, L., D. Wang and W. D. Doye. IEEE Trans. Magn. MAG-31: 2892
(1995)
Heim, D. E., R. E. Fontana, Ja. C. Tsang, V. S. Speriosu, B. A. Gurney
and M. L. Williams. Design and operation of spin valve sensors. IEEE
Trans. Magn. 30: 316(1994)
Hermann, T. M., W. C. Black and S. Hui. Magnetically coupled linear
isolator. IEEE Trans. Magn. 33: 4029 (1997)
Jarratt, J. D., T. J. Klemmer and J. A. Barnard. Microstructural studies of
sputtered C090 Fe,o/Ag GMR multilayers. Mat. Res. Soc. Symp. Proc.
475: 351 (1997)
Johnson, K. E. Proceeding of Diskcon '97 International Technical
Conference. San Jose, Cal ifornia, U. S. A. (1997)
Julliere, M. Tunneling between ferromagnetic films. Phys. Lett. A 54: 225
(1975)
Katada, H. T. Shimatsu I. Watanabe, H. Muraoka, Y. Sugita and Y. Nakamura.
Induced uniaxial magnetic anisotropy field in very thin NiFe and CoZrNb
films. IEEE Trans. Magn. MAG-36: 2905 (2000)
Kawahito, S., Y. Sasaki, H. Sato, T. Nakamura and Y. Tadokoro. A
fluxgate magnetic sensor with micro-solenoids and electroplated permalloy
cores. Sens. Actuators A. A 43: 128 - 134 ( 1994)
Kools, J.C.S. IEEE Trans. Magn. MAG-32: 3165(1996)
Korenivski, V. and R. B. van Dover. Magnetic films for GHz applications. J.
Appl. Phys. 81: 4878 ( 1997)
Lambeth, D. N., E. M. T. Velu, G. H. Bellesis, L. L. Lee and D. E.
Laughlin. J. Appl. Phys. 79: 4496 ( 1996)
Leal, J. L. and M. H. Kryder. J. Appl. Phys. 83: 3720(1998)
Lederman, M. IEEE Trans. Magn. MAG-35: 794 (1999)
Lee, W.-Y., M.F. Toney, P. Tameerug, E. Allen and D. Mauri. J. Appl.
Phys. 87 (9): 6992(2000)
Lenz, J. E. A review of magnetic sensors. Proceedings of the IEEE. 78: 973
(1990)
Liu, Y., B. W. Robertson, Z. S. Shan, S. S. Malhotra, M. J. Yu, S. K.
Renukunta, S. H. Liou and D. J. Sellmyer. IEEE Trans. Magn. MAG-30:
4035(1994)
Liu, Y., Z. S. Shan and D. J. Sellmyer.IEEE Trans. Magn. MAG-32: 3614
(1996)
Lu, P. L. and S. H. Charap. IEEE Trans. Magn. MAG-30: 4230(1994)
Maissel, L.I. and R. Giang. Handbook of Thin Film Technology. McGraw-Hili
Publishing Company, 13-11(1983)
McGuire, T. R. and R. I. Potter. Anisotropic magnetoresistance in
ferromagnetic 3d alloys. IEEE Trans. Magn. MAG-ll (4): 1018 - 1038
(1975)
Miyazaki, T. and N. Tezuka. Giant magnetic tunneling effect in Fe/Ab03/Fe
junction. J. Magn. Magn. Mater. 139: L231 (1995)
336 Dexin Wang

Monsma, D. J. , J. C. Lodder, T. J. A. Popma and B. Dieny. Perpendicular


hot electron spin-valve effect in a new magnetic field sensor: The spin-valve
transistor. Phys. Rev. Lett. 74: 5260(1995)
Moodera, J. S., R. Meservey and P. M. Tedrow. Artificial tunnel barriers
produced by cryogenically deposited A1 2 0 3 . Appl. Phys. Lett. 41: 488
(1982)
Moodera, J. S., L. R. Kinder, T. M. Wong and R. Meservey. Large
magnetoresistance at room temperature in ferromagnetic thin film tunnel
junctions. Phys. Rev. Lett. 74: 3273(1995)
Nakashio, E, J. Sugawara, S. Onoe, and S. Kumagai. IEEE Trans. Magn.
36: 2812(2000)
Neel, L. Memoires et communications. C. R. Acad. Sci. 255: 1545, 1676
(1962)
Parkin, S. S. P., Z. G. Li and D. J. Smith. Giant magnetoresistance in
antiferromagnetic Co/Cu multilayers. Appl. Phys. Lett. 58: 2710C1991a)
Parkin,S. S. P., R. Bhadra and K. P. Roche. Oscillatory magnetic exchange
coupling through thin copper layers. Phys. Rev. Lett. 66: 2152 C1991 b)
Pohm, A. V. , J. M. Anderson, R. S. Beech, and J. M. Daughton. Exchange
coupling and edge pinning in vertical head sensors. IEEE Trans. Magn. 37:
1681(2001)
Pollock, N. Electronic compass using a fluxgate sensor. Wirless World, 1982,
pp.49-54
Primdahl, F. The fluxgate magnetometer. J. Phys. E.: Sci. Instrum. 12: 241-
253(1979)
Prinz, G. A. Spin-polarized transport. Physics Today 4: 58(1995)
Qian, Z. , M. Tondra, D. Wang, J. Daughton, D. Brownell, C. Nordman, and
J. Schuetz. Exsitu CrPtMn pinning and exsitu pinned spin valves. J. Appl.
Phys. 89: 6594 C200 1)
Ripka, P. Review of magnetic fluxgate sensor. Sens. Actuatoers A. A 33:
129-134(1992)
Russek, S. E., J. O. Oti, S. Kata and E. Y. Chen. High-speed
characterization of submicrometer giant magnetoresistive devices. J. Appl.
Phys. 85: 4773(1999)
Sato, M. and K. Kobayashi. Spin-valve-like properties and annealing effect
in ferromagnetic tunnel junctions. IEEE Trans. Magn. 33: 3553 C1997a)
Sato, T., E. Komai, K. Yamasawa, T. Hatanai and A. Makino. Applications
of nanocrystalline Fe Cor Co-Fe) -Hf-O magnetic films for high electrical
resistivity to micro DC-DC coverters. IEEE Trans. Magn. 33: 3310C1997b)
Sato , M. ,H. Kikuchi and K. Kobayashi. Ferromagnetic tunnel junctions with
plasma-oxidized AI barriers and their annealing effects. J. Appl. Phys. 83:
6691 - 6693 C1998a)
Sato, T. , T. Miura, S. Matsumura and K. Yamasawa. New applications of
nanocrystalline FeCCo-Fe)-Hf-O magnetic films to micromagnetic devices.
J. Appl. Phys. 83: 6658C1998b)
Magnetoresistive Thin Film Materials and Their Device Applications 337

Seitz, T. Fluxgate sensor in planar microtechnology. Sens. Actuators A. A 21 - 23 :


799 - 802( 1990)
Sharrock, M. P. J. Appl. Phys. 15: 6413( 1994)
Shen, J.X. and M. Kief. J. Appl. Phys. 79: 5008(1996)
Simonds, J. L. Magnetoelectronics today and tomorrow. Physics Today 4: 27
(1995)
Simmons, J. G. J. Appl. Phys. 34: 1793 ( 1963a) .
Simmons,J. G. Electric tunnel effect between dissimilar electrodes separated
by a thin insulating film. J. Appl. Phys. 34: 2581(1963b)
Smith, C. H. and J. L. Brown. Giant magnetoresistance materials and
integrated magnetic sensors. SAE international congress and exposition
detroit, (1997)
Solin, S. A., T. Tineke, D. R. Hines and J. J. Heremans. Enhanced room-
temperature geometric magnetoresistance in inhomogeneous narrow gap
semiconductors. Science 289: 1530 (2000)
Sousa, R. C et al. Appl. Phys. Lett. 73: 3288 - 3290 ( 1998)
Spoong, J. K., V. S. Speriosu, R. E. Fontana, M. M. Dovek and T. L.
Hylton. Giant magnetoresistive spin valve bridge sensor. IEEE Trans.
Magn. 32: 366(1996)
Stearns, M. B. Simple explanation of tunneling spin-polarization of Fe, Co, Ni
and its alloys. J. Magn. Magn. Mater. 5: 167(1977)
Stoner, E. C. and E. P. Wolhlfarth. Phil. Trans. Roy. Soc. London A 240:
599(1948)
Tehrani, S. , E. Chen, M. Durlam, T. Zhu and H. Goronkin. IEEE IEDM 193
(1996)
Tehrani, S. , E. Chen, M. Durlam, M. DeHerrera, J. M. Slaughter, J. Shi
and V. Kerszykowski. J. Appl. Phys. 85: 5822 ( 1999)
Tehrani, J. , M. Slaughter, E. Chen, M. Durlam, J. Shi and M. DeHerrera.
IEEE Trans. Magn.35: 2814(1999)
Thornton, J. A. J. Vac. Sci. Technol. A 4: 3059 ( 1986)
Tondra, M. , J. M. Daughton, D. Wang, R. Beech, A. Fink and J. Taylor. J.
Appl. Phys. 83: 6688(1998)
Vavara, W. , S. F. Cheng, A. Fink, J. J. Krebs and G. A. Prinz. Appl.
Phys. Lett. 66: 2579-2581(1995)
Wang, D., J. M. Daughton, C. H. Smith and E. Y. Chen. Effect of Au
underlayers on GMR properties of NiFe/Cu/CoFe sandwiches. IEEE Trans.
Magn. MAG-32: 4728(1996)
Wang, D. , J. Anderson and J. M. Daughton. Thermally stable, low saturation
field, low hysteresis, high GMR CoFe/Cu multilayers. IEEE Trans. Magn.
33: 3520( 1997)
Wang, D., J. M. Daughton, K. Bussmann and G. A. Prinz. Magnetic
properties of very thin single and multilayer NiFeCo and CoFe films
deposited by sputtering. J. Appl. Phys. 83(10: 7034(1998)
Wang, D. , M. Tondra, C. Nordman and J. M. Daughton. Thermal stability of
338 Dexin Wang

spin dependent tunneling junctions pinned with IrMn. IEEE Trans. Magn. 35:
2886 ( 1999a)
Wang, D. , M. Tondra, J. M. Daughton, C. Nordman and A. V. Pohm. Spin
dependent tunnel/spin valve devices with different pinning structures made
by photolithography. J. Appl. Phys. 85: 5255(1999b)
Wang, D., J. M. Daughton, D. Reed, W. D. Wang and J. Q. Wang.
Magnetostatic coupling in spin dependent tunnel junctions. IEEE Trans.
Magn. 36: 2802(2000)
Wang, D., M. Tondra, A. V. Pohm, C. Nordman, J. Anderson, J. M.
Daughton and W. C. Black. Spin dependent tunneling devices fabricated for
MRAM applications using latching mode. J. Appl. Phys. 87: 6385(2000b)
Wang, D., M. Tomdra, C. Nordman, Z. Qian, S. M. Daughton, E. Lange, D.
Brownell, L. Tran and Schuetz. Prototype SDT isolators integrated with IC
electronics. GD-07, 46th MMM, Seattle (2001a)
Wang, D. , Z. Qian, J. M. Daughton, C. Nordman, M. Tondra, D. Reed and
D. Brownell. Fabrication and properties of spin dependent tunneling
junctions with CoFeHfO as free layers. J. Appl. Phys. 89: 6754(2001b)
Wang, D., Z. Qian, J. M. Daughton and R. T. Fayfield, US Patent Pending
09/439.892
Wold, R. J., C. A. Nordman,E. Lavely,M. Tondra,E. Lange and M. Prouty.
Development of a handheld mine detection system using a magnetoresistive
sensor array. Proceedings of SPIE vol. 3710, Orlando (1999)
Yaoi, T., S. Ishio and T. Miyazaki. Dependence of magnetoresistance on
temperature and applied voltage in a 82NiFe/AI- AI 2 03/CO junction. J.
Magn. Magn. Mater. 126: 430(1993a)
Yaoi, T., S. Ishio and T. Miyazaki. Magnetoresistance in 82NiFe/AI-AI2 0 3/Co
junction-dependence of the tunneling conductance on the angle between the
magnetizations of two ferromagnetic layers. IEEE Transl. J. Magn. Japan
8: 498(1993b)
Zhu, J. G. and Y. Zheng. IEEE Trans. Magn. 34: 1063(1998)
11 Nano-Structural Magnetoelastic Materials for
Sensor Applications

Keat G. Ong and Craig A. Grimes

11. 1 Introduction

Amorphous ferromagnetic alloys, commonly known as metallic glasses, have


seen considerable interest from sensor researchers since their discovery in the
1960s (Duwez and Lin, 1967; Mader and Nowick, 1965). Metallic glasses are
non-crystallized ferromagnetic alloys free from defects such as grain
boundaries, precipitates, or phase segregations. The unique structure of these
materials gives rise to unique magnetic and mechanical properties, such as
large magnetoelastic coupling, large magnetostriction (Lachowicz and
Szymczak, 1984; Lacheisserie, 1982; Inoue, 2001; Sanchez et aI., 1999;
Boubcheur et al. , 2001), and the so-called magneto-impedance effect (GMI)
(Knobel et aI., 1997a; Ovari et aI., 2001), which make them suitable for
various sensor applications. The large magnetostriction allow the permeability
of the materials to significantly change with applied stress, and also causes the
metallic glasses to exhibit a strong mechanical vibration when excited by a
time-varying magnetic field. The high values of magnetoelastic coupling allow
the metallic glasses to efficiently convert mechanical energy to magnetic
energy (0' Handley, 2000), so the response of the metallic glasses can be
remotely detected by measuring the magnetic flux change. Furthermore, the
GMI effect, a phenomenon that significantly changes the electrical impedance
of the metallic glasses with a small variation in applied magnetic field, allows
them to be used as ultra-sensitive magnetic field sensors. Most metallic
glasses have excellent soft magnetic properties such as low coercivity, low
hysteresis loss, and high permeability, due to the fact that there are no
defects restricting the magnetic domain wall motion. The soft magnetic
properties increase the performance and the stability of metallic glass-based
sensors by allowing them to have good reversibility with little hysteresis in
response to applied mechanical stress and magnetic fields.
Metallic glasses are fabricated by rapid quenching of a molten alloy to
prevent crystallization. The properties of the materials are largely dependent
upon the alloy composition, which is generally a combination of transition
metals, rare-earth metals in the case of magnetically hard alloys, and glass-
340 Keat G. Ong and Craig A. Grimes

forming metalloids. Inhomogeneous cooling of the alloy melt creates internal


stresses that increase the anisotropy field. However the amorphous nano-
structure of these materials allows atomic rearrangement at a temperature far
below crystall ization, thus the stress can easily be relaxed with low-field
annealing at a moderate temperature. Alternatively, depending upon the
application, an anisotropy field can be introduced in a metallic glass by
annealing under applied stress or magnetic field at moderate temperatures
(Brouha and van de Borst, 1979; Soyka et aI., 1999; Price and Overshott,
1983; Modzelewski et al. , 1981).
Although there are many types of metallic glasses with a range of
mechanical and magnetic properties depending upon the composition and
annealing conditions, the most useful type for sensor applications is the
magnetically soft transition-metalloid-based alloy with a high content of Fe or
Co; Fe-rich alloys have magnetostriction constants as high as }. s _10- 5 , and
magnetoelastic couplings as high as O. 98 (Gutierrez et aI., 1989; Chiriac
et al. , 2000a; Lanotte et al., 2000; Mitchell et al., 1979), while Co-rich
transition-metalloid alloys have low magnetostriction in the order of 10- 7
Some of the commercially available metallic glasses for sensor applications
are the Fe-rich Metglas 2826 MB (Fe4o Nh8 Mo4B18 ) and 2605SC (Feal B13 . 5 Si3.5 G.!)
(Spano et aJ., 1982; Savage and Spano, 1982; Copeland et aI., 1994)
ribbons fabricated by Honeywell International Inc. CD, which have a
magnetostriction coefficient in the order of 105 and a high magnetoelastic
coupling coefficient.
Metallic glasses are commonly chosen as the core materials of induction
sensors for the measurement of physical properties such as stress/strain and
torque (Spano et aJ., 1982; Hernando et aJ., 1988), magnetic field, and
displacement (Mohri, 1984) due to their changing permeability with applied
stress and field. These sensors, usually composed of solenoids loaded with
metallic glass cores, are simple in design, stable over a wide range of
temperatures, and have a good linearity and low hysteresis.
Metallic glasses also exhibit excellent GMI effects at radio frequency and
are extensively used for the detection of magnetic fields and stress. GMI
sensors based on metallic glasses are highly sensitive, with maximum
impedance variation reaching 400 % (Bushida et aI., 1995, 1996; Kitoh
et aI., 1995; Mohri et aJ., 2001). Metall ic glasses have been used to
construct magnetostrictive delay lines (MOL) for the measurement of stress,
torque, and displacement (Chiriac et aJ. , 1997a, 2000b, 2001 a; Hristoforou
et al., 1992, 1996, 1998, 1999; Germano et al., 1997, 2000). The MDL
sensor is usually a long metallic glass wire or ribbon surrounded with an
excitation coil at one end and a sensing coil at the opposite end. An acoustic

CD Honewell,101 Columbia Road, Morristown, NJ 07962 USA. http://www. honywell.


com.
Nano-Structural Magnetoelastic Materials for Sensor Applications 341

wave is excited by the excitation coil with a voltage pulse, and is captured by
a sensing coil as an induced voltage. The parameters of interest, such as
stress, torque, and weight, are determined from the attenuation of the
acoustic wave, which is obtained by normalizing the measured voltage
amplitude to the excitation voltage amplitude.
Magnetoelastic resonance sensors, which are based on the change in the
magnetoelastic resonant frequency of a metallic glass wire or ribbon towards
applied stress and field, have become popular in the past decade because of
their remote-query capability, simple design, and low cost (Gutierrez and
Barandiaran, 1995; Grimes et al. , 1999a). These sensors have been used to
measure pressure (Grimes and Kouzoudis, 2000a), liquid viscosity and
density (Jain et al., 2000), liquid flow velocity (Kouzoudis and Grimes,
2000b), temperature (Mungle, 2001), and thin-film elasticity (Schmidt,
2000). Magnetoelastic resonance chemical and gas sensors have also been
built by applying a layer of mass-changing chemically responsive material, the
mass loading of which changes the mechanical resonant frequency of the
sensor, as does the elasticity of the appl ied coating (Grimes et aI., 1999a,
1999b, 1999c, 2000c, 2000d).
Metallic glass thin films are also used as an active layer of a magnetic
surface acoustic wave (MSAW) sensor (Chiriac et al. , 2001b). Due to the
stress and field dependencies in the metallic glass, the MSAW sensor can be
used to measure stress, magnetic field, or displacement. Recently, a
viscosity sensor (Vazquez et al. , 2001) was built based on the spontaneous
mechanical rotation of a large magnetostrictive metallic glass wire when
subjected to an ac magnetic field.

11. 2 Metallic Glass Preparation and Properties

This section includes the preparation process, properties, and methods of


enhancing the properties of the metall ic glass alloys. Section 11. 2. 1 is the
discussion of the preparation process and alloy compositions of the metallic
glasses. Section 11. 2. 2 presents the physical and magnetic properties of
metallic glasses, detailing how metallic glasses exhibit magnetostriction,
magneto-impedance, and magnetoelastic resonance. Section 11 .2.3
describes the ways to change the properties of metallic glasses by varying
alloy composition and annealing.

11.2.1 Preparation of Metallic Glasses


Metallic glasses are commonly fabricated with melt quenching methods,
342 Keat G. Ong and Craig A. Grimes

including melt-spinning (Lieberman and Graham, 1976), electrodeposition


(Fukamichi et aI., 1997), sputtering (Lu and Nathan, 1997; Donton et al. ,
2000), and ion plating (Uchida et aI., 1994). The key to successfully
fabricating a metallic glass is rapid removal of heat from the melt, to preclude
crystallization in the alloy. In most cases, cooling rates of 105 'c /s are
needed. Most of the magnetically soft metallic glasses are based on 3d
transition metals (T) with glass forming metalloids (M) such as boron,
carbon, silicon, or phosphors to stabilize the amorphous state. The
composition of transition-metalloid-based alloys, T 1 - x Mx ' is usually in the
range of 15 < x < 30 at. % , such as Feao B20 ' Fe40 Ni40 P14 B6 , and ~4 Fes B18 Sh .
The late transition metals (TL = Fe, Co, Ni) can also be stabilized in an
amorphous state by alloying with early transition metals (Zr, Nb, Hf), forming
TE 1- x TL x ' with x in the range 5 at. % - 15 at. % such as C090 Zrl0' Among
these metallic glasses, the transition-metalloid based alloys have the best
properties for sensor applications, such as a high magnetostrictive and
magnetoelastic coupling, and a low anisotropy and hysteresis.

11. 2. 2 Summary of Physical and Magnetic Properties of


Metallic Glasses
Metallic glasses are flexible, with Young's moduli of around 100 GPa, about
20 % - 30 % lower than their crystalline counterparts (Hernando et aI., 1988;
Knuyt et aI., 1986). Metallic glasses also exhibit high mechanical strength
and ductility. For example, Metglas 2605SC and 2826 MB ribbons have a
tensile strength of 1000 - 1700 MPa. The electrical resistivity of metallic
glasses are generally one order of magnitude higher than their crystalline
counterparts, with typical values of 100 - 200 IJQ cm (Copeland et al. ,
1994). In general, the metallic glasses, especially those made of transition
metals and metalloids, are magnetically soft due to high electrical resistivity
and low anisotropy. Common ferromagnetic metallic glasses have magnetic
susceptibilities on the order of 105 , saturation magnetizations of about 1 T
(Hernando et aI., 1988), and compared to their crystalline counterparts, low
hysteresis losses.
Note that the values given above are those of common materials, as
generally fabricated and used; properties can be altered by changing alloy
composition and performing annealing, as will be detailed in Section 11.2.3.
In addition to these properties, metallic glasses also exhibit unique behaviors
such as magnetoelastic and magnetostriction effects, magnetoelastic
resonance, and giant-magneto-impedance effect, which will be explained in
the following sub-sections.

11.2.2.1 Magnetoelastic and Magnetostriction Effects


Most metallic glasses exhibit magnetostriction, a phenomenon where the
Nano-Structural Magnetoelastic Materials for Sensor Applications 343

dimensions of a magnetic object change when subjected to the influence of an


external magnetic field. The magnetostriction of a ferromagnetic glass varies
widely with alloy composition, and depends upon applied stress, structural
relaxation, and fabrication parameters (Brouha and van de Brost, 1979;
Soyka et aI., 1999; Price and Overshott, 1983; Modzelewski et aI., 1981).
Generally speaking, Fe-rich alloys exhibit high magnetostriction values, A s ~
10- 5 , while Co-rich alloys have magnetostrictions of approximately As _10- 7
(Hernando et aI., 1988).
A model was developed to understand the effects of applied stress and
applied magnetic field on susceptibility, magnetoelastic coupling, and
Young's modulus of a ferromagnetic metallic glass ribbon (Livingston, 1982).
The model, shown in Fig. 11. 1, considers a ribbon-shaped metallic glass
having a width-wise magnetic easy axis. The anisotropy of the ribbon is
assumed uniaxial with an energy contribution of K u cos e per unit volume.
With zero applied field and zero applied stress, the magnetization in the
domains will align across the width of the ribbon. With a longitudinal applied
field H and a longitudinal stress 0, the magnetization rotates from the width
towards the length direction (Fig. 11. 1), yielding a magnetostrictive strain E
given by (Livingston, 1982):

E
= EM
..!!...- + 3 As (H
2

2 Hz,.
_ l-)
3
(H < Hk ) (11. 1)

where EM is Young's modulus at a constant magnetization, and the anisotropy


field H k is (Livingston, 1982):
_2Ku-3AsO
Hk - M . (11. 2)
s
The magnetic susceptibility X is related to the applied strain as
_ M~
(11.3)
X - 2K u -3AsO
where M s is the saturation magnetization. The sensitivity of the susceptibility
is given by (Hernando et aI., 1988):

dX
do
= (2K M
2
3A s. (11.4)
u - ;Aso)
It is clear that the sensitivity of the susceptibility increases with higher As and
lower K u. The magnetoelastic coupling factor k can be related to the
parameters already derived as

k =
MsH~
( 1 + 9EMA~H2
)-2 1

(11. 5)

Equation (11. 5) shows that a high magnetoelastic coupl ing requires a high As
and EM' and a low H k' For applications based on changes in the elastic
modulus, for example a strain sensor, the parameter of interest is the
344 Keat G. Ong and Craig A. Grimes

fractional difference between the elastic modulus at constant magnetization EM


and the modulus at constant field E H' or the so-called l:!. E effect, which is
given by (Livingston, 1982):

l:!. E _ EM - E H _ 9A~ E MH
2
_ k
2
(H < H ) ( 11 .6)
E-,; - EH - MsH~ - 1- k 2 k .

Equation (11.6) indicates that the condition of a high l:!.E effect is similar to the
magnetoelastic coupling.

Figure 11.1 Magnetic domain of a transverse-field annealed metallic glass ribbon at (a)
zero bias field (H=O) and (b) O<H<H k

11.2.2.2 Magnetoelastic Resonance


A ribbon- or wire-shaped metallic glass exhibits magnetoelastic resonance
when excited by a time-varying magnetic field. For a metallic glass ribbon
with its length parallel to the x -axis of a rectangular coordinate system,
excited by an x-directed ac magnetic field, the ribbon exhibits a vibration that
can be described with the equation of motion as (Grimes et aI., 1999a):
aUx EH a2 U x (11. 7)
Ps at2 = 1 - O"~ ax 2

where U x is the displacement in the x -direction, Ps is the density of the


metallic glass ribbon, and O"p is the Poisson's ratio. Although the ribbon
exhibits vibrations at almost every frequency, the vibrations are most
pronounced at the resonant frequencies (Grimes et al. , 1999a):

f = -.!2..- / EH (n = 1,2,3,) (11.8)


n 2L,yPs(1-0"~)
where L is the length of the ribbon. For most sensor applications, only the
first harmonic (n = 1) is considered due to its high amplitude. The frequency
response of an 80 mm x 3 mm x 30\Jm as-cast Metglas 2826 MB ribbon is
plotted in Fig. 11. 2, where the resonant frequency f o is the maximum
amplitude of the frequency spectrum and the anti-resonant frequency fa is the
minimum.
The effective magnetoelastic coupling k can be related to f o and fa as
(Anderson, 1982; Mithchell et ai., 1979)

k2 = 1 _ f;f~ . (11.9)

The intrinsic magnetoelastic coupling of the material, k 33 , which is geometry


Nano-Structural Magnetoelastic Materials for Sensor Applications 345

0.12
10
0.1

~ 0.08
<l)
bJl
;S
0.06
~
0.04

0.02
28.6 28.8 29 29.2 29.4
Frequency (kHz)

Figure 11. 2 The frequency response of an 80 mm x 3 mm x 30 ~m as-cast fv1etglas 2826 MB.

independent, is given by (Modzelewski et al. , 1981)

_f~)
2
2 _Tr (
(11.10)
k 33 - 8 1 f;'
Many authors (Gutierrez et aI., 1989; Anderson, 1982; Modzelewski et al. ,
1981) used k 33 instead of k to describe the magnetoelastic coupling of the
metall ic glasses due to its geometry independence.
11.2.2.3 Giant Magneto-Impedance Effect
Giant magneto-impedance (GMI) is a phenomenon that significantly changes
the high frequency electrical impedance of a magnetic conductor with a small
variation of magnetic field. Metallic glasses, especially Co-based glasses with
near zero magnetostriction ('" 10- 7 ) , display excellent GMI effect, which
makes them suitable for the measurement of magnetic fields in the range of few
Oersted. To measure the GMI effect, generally a metallic glass wire is
connected to an ac current source and voltmeter, and the impedance is
determined from the ratio of the measured voltage to the applied current. The
sensitivity of the GMI in a metallic glass is largely dependent upon the
frequency of the driving current and the magnitude of the transverse
permeability in the glass. For a ribbon shape metallic glass, the impedance
measured across the length is given by (Machado and Rezende, 1996):
(1-j)L 1
Z = 2 (2Trpe w!Jt)"2 (11.11)
we
where j is the complex number, e is the speed of light, wand L are
respectively the width and length of the ribbon, Pe is the electrical resistivity
of the metallic glass, W is the radian frequency of the driving current, and !JI is
the transverse permeability of the metallic glass. For a circular metallic glass
wire, the impedance measured across the length is (Bushida et aI., 1995)
346 Keat G. Ong and Craig A. Grimes

el1.12)

where a is the radius of the wire, R de is the dc resistance, and IJ e is the


circumferential linearized permeability.
The GMI effect is usually expressed as a function of applied field as the
GMI ratio, GMI %, which is defined as

GMI% = ZeHapp,)- ZeH max ) ell.13)


ZeH max )

where Z is the impedance, and Happ, and H max are respectively the applied and
saturation magnetic fields. The GMI% for metallic glasses are relatively high,
normally in the range between 100% - 500%, depending upon the metallic
glass composition, heat treatment, bias dc current, and the design of the
sensor.

11. 2. 3 Tailoring of Metallic Glasses for Sensor Applications


This section describes the methods to control the mechanical and magnetic
properties of metallic glasses. Section 11.2.3. 1 explains how alloy
compositions have a direct impact on the magnetoelastic properties of
transition-metal-based metallic glasses, while Section 11.2.3.2 describes the
method of increasing the magnetoelastic properties of metallic glasses via
transverse-field annealing and stress annealing. Section 11.2.3.3 explains the
enhancement of GMI effect in metallic glasses through low field annealing,
increasing the operation frequency, and applying external stresses.
11. 2. 3.1 Effects of Alloy Compositions on the Metallic Glass Properties
Alloy compositions playa direct role in the magneto-mechanical properties of
the metallic glasses such as magnetostrictive and magnetoelastic couplings, as
well as the magnetic properties such as magnetic hardness, hysteresis losses,
and anisotropy. Fe-rich transition-metalloid-based alloys can have substantial
magnetostriction constants and high magnetoelastic couplings because they
have low anisotropy energy K u and low anisotropy field H k' which permit the
rotation of the magnetization and accompanying magnetostrictive strain with
low applied fields or stresses. For example, Metglas 2605SC, which has a
high magnetostriction of 3 x 10- 5 and magnetoelastic coupling of 0.98 has a
low K u and H k of 38 Jim and 370 Aim, respectively.
It is found that the induced anisotropy constant K u of Fe-rich alloys
increases strongly with the presence of additional transition elements such as
Ni or Co eLivingston, 1982). This is the reason why the magnetoelastic
coupling of the 2826 M8 Fe40 Nbs M048 1S ek ~O. 5) is inferior compared to the
2605SC FeSl 8 13 . 5 Sb.5Cz ek = 0.98).
Nano-Structural Magnetoelastic Materials for Sensor Applications 347

11. 2. 3. 2 Increasing Magnetoelastic Coupling via Annealing


The model in Section 11. 2. 2. 1 indicates that a width-wise magnetic domain
distribution and a low uniform H k are required for achieving a high
magnetoelastic coupling k. However, in an as-cast metall ic glass ribbon, the
surface defects and irregularities create localized stresses and cause an
inhomogeneous K u distribution on the ribbon, leading to variations in the
direction and amplitude of localized magnetic domains and H k' and in turn
reduce the magnetoelastic coupling. It is thus necessary to remove the internal
stresses and orient the magnetic domains width-wise without including
significant crystallization.
Transverse field annealing is the most common method used to remove
internal stresses for increasing ih and k of ferromagnetic glasses. Extensive
investigations have been conducted to analyze the effects of annealing
temperature and time on As, k, and the magnetic permeability IJ of various
metallic glasses. For example, Anderson (1982) found the magnetoelastic
coupling of a Metglas 2605SC ribbon can reach a value higher than O. 9 at a
bias field of '" O. 6 Oe after annealing at 405 - 415 "C. He also found the
maximum !:l.E effect occurs at the maximum k for 2605SC. Modzelewski et al.
(1981) also performed annealing studies for Metglas 2605SC, and the results
are similar to Anderson's. Mitchell et al. (1979) examined the effects of
annealing Fe71 CoB 20 over various annealing temperatures and times, and found
that optimal k can be obtained at 350 - 390C for 5 - 30 min. Mitchell et al.
also noticed k falls off rapidly as the annealing temperature approaches the
crystall ine temperature at about 400 C. Price and Overshott (1983) measured
the initial permeability of Metglas 2826MB after annealing under temperature
between 325 and 425 "C , and found the highest permeability can be achieved
at 345C .
Alternatively, annealing under stress and field can also lower H k .
According to Eq. (11.2), H k is proportional to the term (K u - 3/2 i\ sa) , thus
by choosing an appropriate stress a one can set the term to zero and obtain a
low H Ii , Stress annealing studies were conducted by Nielsen (1985), Vazquez
et al. (1987), and Gonzalez et al. (1987), and showed that annealing under
stress and field gives rise to an unexpected increase in the magnetoelastic
coupling. Recently Soyka et al. (1999) annealed amorphous Feso Cr2 B 14 Si 4
ribbons under longitudinal tensile stress and transverse fields (H = 400 kA/m) ,
and found that the !:l.E effect is almost constant and large at tensile stress less
than 1OOMPa, but rapidly decreases at higher stress levels. He also found that
Young's modulus of the ribbon decreases much faster at low bias fields for a
low stressed sample.
11.2.3.3 Methods of Increasing the GMI Effect
Equations (11. 10 and (11. 12) show the impedance of the metallic glass
increases with the factor J WlJt.8 for both metallic glass wires and ribbons.
348 Keat G. Ong and Craig A. Grimes

Based on this fact. one can increase the sensitivity of a GMI sensor by
increasing the operating frequency. In addition. it is also found that the
transverse and circumferential permeabilities are dependent upon the
amplitude of the driving current. applied stress. and annealing temperature.
The effect of ac frequency on the GMI % (giant-magneto-impedance ratio)
was studied by Knobel et al. (1997 a) and Ovari et al. (2001). and they found
the GMI % increases with frequency. but decreases when exceeding a certain
threshold frequency. The threshold frequency. generally around 1 - 200 MHz.
is dependent upon the types of metallic glasses and annealing conditions. The
amplitudes of the dc bias and ac currents, and the internal stress of the
metallic glass have also found to influence the sensitivity of the GMI % (Allia
et al. , 2001; Kawashima et al. , 1999). However, these parameters interact
with each other and also depend on the alloy composition of the metallic glass.
and there are yet no conclusive observations as to how they affect the GMI % .
Annealing of metallic glasses can improve the sensitivity of GMI%, and
the degree of improvement is directly related to the annealing temperature. An
extensive investigation on the annealing effect on the GMI % has been
conducted by Takemura et al. (1996), Knobel et al. (1997b), AIIia et al.
(2001), Moya et al. (1999) and He et al. (2001). All of them found that
increasing annealing temperature increases the GMI %. However, they also found
the GMI% drops as the annealing temperature reaches the crystallization
temperature of the metall ic glass, which is around 500 - 600 "C. Recently, Brunetti
et al. (2001) improved the sensitivity of the GMI sensors by 600% by coating a
glass layer on a Fe73.5 Cu3 Nb1 Si n . 5 8g metallic glass microwire. Cobeno et al.
(2001) also altered the sensitivity of the GMI % by applying a tensile stress on a
glass-coated Cot;S.5 Mn6.5 SilO B15 metallic glass microwire, and they found the
maximum improvement for the GMI% occurs at around 60 MPa.

11. 3 Sensor Applications

Section 11. 3. 1 shows the applications of metallic glasses as the magnetic


cores of induction sensors for measuring field, stress and torque, and
displacement, while Section 11.3.2 shows the applications of metallic glass
wires as magneto-impedance sensors. Section 11. 3. 3 illustrates how metall ic
glass wires are used to construct magnetostrictive delay lines for the
measurement of stress, torque, and displacement. Magnetoelastic resonance
sensors, which have been used to measure pressure, liquid flow and density/
viscosity, temperature, elasticity, and chemical concentration, are presented
in Section 11. 3. 4. Sections 11. 3. 5 and 11. 3. 6 demonstrate the uses of
metallic glasses as the sensing elements of MSAW sensors and gyro-magneto-
sensors, respectively.
Nano-Structural Magnetoelastic Materials for Sensor Applications 349

11.3.1 Induction Sensors with Metallic Glass Cores


The induction sensor is composed of a solenoid or a coil with a permeability/
susceptibility-changing metallic glass core. Since the inductance of the coil is
proportional to the susceptibility of its core material, changes in the metallic
glass susceptibility can be determined from the change of the coil inductance.
It is also common to use a pair of excitation and sensing coils, where the
excitation coil excites the metallic glass with an ac magnetic field, and the
susceptibility of the metallic glass is determined from the amplitude of the
induced voltage recorded by the sensing coil.

11.3. 1. 1 Measurement of Stress/Strain


As indicated in Eqs. (11.3) and (11.6), the susceptibility and elasticity of
the metallic glasses are strongly dependent upon the applied stress and field.
As a result, stress/strain are generally monitored by measuring the
susceptibility of the metallic glass with an inductive coil.
As shown in Fig. 11.3, the inverse susceptibility of the metallic glass 1/X
shows good linearity with the applied stress (Spano et aI., 1982). For
metall ic glasses with positive magnetostriction, 1/ X is inversely proportional
to (j until reaching the demagnetization factor N, where 1/ X becomes
independent of (j. For negative i\ s' 1/ X is directly proportional to (j starting
from N. To directly compare the sensitivity of a metallic glass sensor to other
semiconductor-based stress sensors, the figure of merit for the metallic glass
sensors is derived as (Spano et aI., 1982; Hernando et al. , 1988)

(11.14)

As<O

o o

Figure 11.3 The inverse susceptibility of the metallic glass X -1 linearity decreases with
the applied stress when the metallic glass has a positive magnetostriction, but linearly
increases when the metallic glass magnetostriction is negative.
350 Keat G. Ong and Craig A. Grimes

Metglas alloy 2605SC has been widely used for stress sensing due to its
high magnetoelastic coupling coefficient, 0.98 after transverse field anneal ing,
and high magnetostriction of :::::::: 3 X 10- 5. The figure of merit of 2605SC
Metglas sensors has been calculated to be in the range of 105 (Spano et al. ,
1982), three orders of magnitude greater than semiconductor strain gauges.
Different experimental configurations were designed to measure the
susceptibility shift of the metallic glass sensors for tensile, compressive, and
torsion stresses. Figure. 11. 4a is a basic tensile stress sensor designed by
Mitchell et al. (1986) and later improved by Barandiaran and Gutierrez
(1997) by connecting the output to a bridge circuit for eliminating the
background signal and enhancing the sensitivity of the sensor. The
susceptibility of the setup in Fig. 11. 4a has a linear response to tensile stress.
Figure. 11. 4b is an improved toroid-shape sensor designed by Mohri and
Korekoda (1978) and Meydan et al. (1981). The four terminals in the setup in
Fig. 11.4b are connected in a bridge configuration so the signal is zero in an
unstressed case, but positive to tensile stress and negative to compressive
stress. A stress sensor designed for compressive stress is illustrated in
Fig. 11. 4c, where the sensor is stretched by a spring, and the actual stress
experienced by the sensor is reduced when compressive stress is exerted
(Hernando et al. , 1988).

2 2'

a I'
(a) (b) (c)

Figure 11.4 (a) A metallic glass ribbon used to measure tensile stress; (b) A metallic
glass toroid used to measure tensile and compressive stresses; (c) A metallic glass sensor
used to measure compressive stress.

Another application based on metallic glass stress sensors is the shock


detection proposed by Mohri and Takeuchi (1981). The shock sensor is
essentially a stress sensor surrounded by an excitation coil and a sensing coil.
When a shock is applied to the sensor, it causes a sudden change in the
metallic glass permeability, which will be captured by the sensing coil as a
voltage spike.
11.3. 1. 2 Torque Measurement
Figure. 11.5a illustrates a novel torque sensor comprised of two metallic glass
Nano-Structural Magnetoelastic Materials for Sensor Applications 351

strips attached 45 to the shaft surrounded by an excitation coil and two series-
connected sensing coils. In the absence of torque, the symmetrical
configurations of the two metallic glass strips and the two sensing coils
eliminate the background signal. When torque is applied, the metallic glass
strips are stressed or compressed, destroying the symmetry of the magnetic
field pattern and causing an induction voltage at the sensing coils. The output
is a type of sensor that is linear and sensitive, up to 1.4 V/ (N m) at 20 kHz
and has been reported by (Harada et aI., 1982; Sasada et aI., 1984). The
sensitivity of the sensor is also found to be proportional to the frequency.

~etallic glass
Excitat~ CT+~enSing {( shaft

coil ~ ~il CT

Primary yoke Secondary yoke


(a) (b)

Figure 11.5 (a) A coaxial torque sensor consisting of two metallic glass ribbons. In the
absence of torque. the sensing coil receives no signal due to the symmetrical
configurations of the two metallic glasses. When a torque is applied to the shaft, the
metallic glasses are compressed or stretched. destroying the symmetry of the magnetic
field pattern and causing an induction voltage at the sensing coil; (b) A torque sensor
based on a four-arm orthogonal yoke design. where a signal will appear in the secondary
coil if a torque is applied to the metallic glass shaft and an ac source is excited at the
primary coil.

Another metallic glass-based torque sensor is a four-arm orthogonal yoke


design shown in Fig. 11.5b (Hilzinger, 1985). There is no output signal in the
absence of torque due to the symmetrical magnetic field pattern on the shaft.
However. when a torque is appl ied at the ends of the metall ic glass shaft, due
to the changes in the magnetic field pattern, a signal will appear in the sensing
coil at the secondary yoke if an ac field is excited at the primary yoke.
11.3.1. 3 Measurement of Magnetic Flux
Due to their high susceptibility and low hysteresis loss, metallic glasses are
widely used as the cores of magneto-inductance sensors to measure magnetic
flux. Since the major requirements of a flux sensor are stability and
independence from stress. most metallic glass flux sensors are made of Co-
rich amorphous alloys, as they have almost zero-magnetostriction (10- 7 ) .
Moldovanu et al. (1997, 2000) designed an induction-based magnetic flux
sensor by inserting two C068 . 25 Feu Si t 2.25 8 15 metallic glass ribbons inside a
sensing coil and an excitation coil. When an external magnetic field is applied
to the metall ic glasses, it changes the apparent permeab iIity of the metallic
352 Keat G. Ong and Craig A. Grimes

glasses, and in turn changes the voltage received by the sensing coil. The
sensitivity of the sensor is highest when the magnetic field is parallel to the
basal-axis of the coil. The output voltage of the sensor is found linear with
magnetic field with sensitivity between 4. 5 and 5. 5 IJV/nT, and for fields
greater than lOOnT it is independent of temperature variation within 20 - 70 C .
Using the same principle, Nielsen et al. (1990, 1995) also designed a flux
sensor based on Vitrovac 6025 alloys, and Ghatak and Mitra (1992) based on
CoFeCrSiB. The sensitivity of Nielsen's sensor is 5 - 10 IJV/nT and the
sensitivity of Ghatak's sensor is 3 IJV/nT.
11.3.1.4 Displacement Sensors
Metallic glasses are also used to measure the displacement of an object by
attaching a permanent magnet on the object (Hernando et al., 1988). As
shown in Fig. 11. 6a, when the location of the magnet changes, the magnetic
field exerted on the metallic glass also changes. Hence, the displacement of
the object can be determined from the changes of the metallic glass
permeability. Alternatively, the object of interest can be attached to the end
of the ribbon. As the object is moving away from the sensor, it pulls the
metallic glass ribbon and changes the applied stress, in turn altering the ribbon
susceptibility (Fig. 11. 6b). Using this principle, Mohri (1984) came up with a
displacement sensor that can detect a displacement of 1 IJm.

[]JJJJ]
Metallic
glass ! ! Moving object
(Permanent magnet)

(a)

~:::lliC [[[[rUf==Sha====jft D
1 1 Moving object

(b)

Figure 11.6 (a) A displacement sensor uses a permanent magnet as a marker. The
displacement of the magnet changes the dc bias field on the metallic glass, which in turn
changes the permeability; (b) A displacement sensor where the object of interest is
attached to the end of a metallic glass ribbon. As the object moves, it pulls the metallic
glass, creating a stress on the metallic glass and changing the permeability.

11.3. 1. 5 Flow Meter


A flow meter, shown in Fig. 11.7, was designed by Hristoforou et al. (1997).
The flow meter is composed of two metallic glass wires: one has both ends
fixed, the other has only one end fixed. The wire with one end fixed is
interrogated and detected with an excitation coil and a sensing coil. The
Nano-Structural Magnetoelastic Materials for Sensor Applications 353

excitation coil is wound around the outside of the flow conduit wall, while the
sensing is two separate coils connected in series opposition as shown in
Fig. 11. 7. When the liquid flow rate is zero, the wire is hanging vertically,
resulting in a zero output voltage V 1 due to the symmetry of the two series-
connected sensing coils. When the fluid flows, the wire is pushed away from
the flow direction, destroying the symmetrical configuration and causing a
voltage VI to appear at the sensing coils. The voltage V, can be positive or
negative, depending upon the flow direction (Fig. 11. 8a) .
AC current
Metallic glass wire / source

AC current
source

AC voltmeter

Excitation coil
AC voltmeter

Figure 11.7 A flow meter composed of two metallic glass wires. one with both ends fixed
and one with only one end fixed. The voltage measured across the wire with both ends
fixed increases with flow rate due to the increasing stress created by the flowing liquid.
The wire that has only one end fixed is detected using a sensing coil connected in series
opposition. When the fluid flows, the wire is pushed and a voltage is recorded at the
sensing coil.

VI (mV)
400 14

L/h :>
-----j1-::----:60 5
::..N

700
L/h
(a) (b)

Figure 11.8 (a) The voltage measured at the sensing coil, V" where it can be positive
or negative depending upon the flow direction; (b) The voltage across the wire with both
ends fixed, V l , has a parabolic profile with the fluid flow rate.

The wire with both ends fixed is connected to an ac current source and
voltmeter. Due to the stress-impedance effect (refer to Section 11. 3. 2), the
voltage measured by the voltmeter V2 is proportional to the stress applied on
354 Keat G. Ong and Craig A. Grimes

the wire. When the flow rate of the liquid is zero, the amount of the stress on
the sensor is equal to the tensile stress created by the two end fixings. When
laminar fluid flow passes through the wire, a parabolic profile of force is
created along the length of the wire, resulting in an additional stress on the
wire and an induced voltage V2 (Fig. 11. 8b).

11.3.2 Giant Magneto-Impedance / Stress-Impedance Sensors


The basic design of a GMI sensor is illustrated in Fig. 11.9. The two ends of a
metallic glass ribbon are connected to a dc-biased ac current source and a
voltmeter, and the impedance of the metallic glass Z is determined by taking
the ratio of the measured voltage to the applied current. A typical response of
a GMI sensor is illustrated in Fig. 11. 10 (Kurlyandskaya et al. , 2001), where
it shows the GMI % increases rapidly with applied magnetic field H when H is
smaller than H k ' but decreases with H for H > H k

~--o
Eo"!.0-------,

Figure 11.9 A magneto-impedance sensor. The two ends of a metallic glass ribbon are
connected to a dc-biased ac current source and the voltage across the two ends is
measured.

Figure 11. 10 A typical response of a magneto-impedance sensor. The GMI ratio


increases rapidly at low magnetic field H until the anisotropy of the metallic glass, where
the GMI ratio starts falling off with increasing H.

11.3.2. 1 Giant Magneto-Impedance Sensors Based on Metallic Glasses


Utilizing the GMI effect, metallic glass wires are widely used as
Nano-Structural Magnetoelastic Materials for Sensor Applications 355

magnetometers for detecting magnetic field in the range of few Oersted. Some
commonly used metallic glass alloys for GMI sensors are Fe4.35 COS8 .15 Si 12 .5B 15
and C072.5Si12.5 B'5 (Kitoh et aI., 1995), COS8 .5 MnS.5 SilO B'5 (Cobeno et al. ,
2001), and Fe73.5Cu,Nb3Si,3.5B9(Knobel et aI., 1997b).
To increase the sensitivity, linearity, and signal-to-noise ratio of the GMI
sensors, usually the metallic glass wires are connected to oscillator circuits,
such as the Colpitts oscillator (Bushida et aI., 1995; Uchiyama et al. , 1995)
and the double Hartley oscillator (Bushida et al. , 1996). Figure 11.11 (a) is a
Colpitts oscillator that utilizes the resonance of the inductance of the metallic
glass and the capacitance C, and C2 to create an ac signal to excite the
metallic glass (Uchiyama et aI., 1995). This design allows the GMI% of the
metallic glass to increase several times at frequencies of~100 MHz. Since the
oscillator output roughly follows the GMI curve, this sensor design provides
non-linear results similar to the curve shown in Fig. 11. 10. By applying a
longitudinal bias field, Bushida et al. (1995) were able to shift the curve in
Fig. 11. 10 horizontally and allowed the metallic glass to operate at linear
region within the range of interest.

+-----+--+--8>I-------,--OUI

(a) (b)

Figure 11.11 A magneto-impedance sensor connected to (a) a Colpitts oscillator. and


(b) a double Hartley oscillator. Although more complicated in design. the double Hartley
oscillator provides better linearity in output voltage.

An oscillator circuit that provides linear output, the double Hartley


oscillator, is shown in Fig. 11. 11 b (Mohri, 1997; Inada et al. , 1994; Bushida
et aI., 1996). For this design, two identical metallic glasses biased by two
opposing magnetic fields H b are needed. As shown in Fig. 11. 12, the output
of this sensor is almost linear in the range of - H b < H < H b This type of
sensor also provides higher sensitivity. For example, in Bushida' s design
(Bushida et al. , 1996), the sensitivity of the GMI % can reach to lO- s Oe at
1 Oe full scale for an ac field. The biasing fields H b , however, require small
solenoids wound around the metallic glasses, which is a major restriction for
sensor miniaturization. By using twisted amorphous wires with dc bias
currents, K itoh et al. (1995) were able to avoid using solenoids and yet retain
the linear characteristic for a limited range of magnetic field.
356 Keat G. Ong and Craig A. Grimes

Figure 11.12 The output of the double Hartley oscillator. where E out is linear when the
applied field is smaller than the bias field.

11. 3.2.2 Stress-Impedance Sensors Based on Metallic Glasses


When a tensile stress is applied to a near-zero magnetostriction metallic glass.
it changes the saturation magnetostriction coefficient. causing a variation in
the metallic glass permeability, and in turn, the impedance. Utilizing this
stress-impedance (Sl) effect, a stress sensor was constructed by passing a
high-frequency current through a C066 . 3 Fe3.7 Si 12 B 18 metallic glass ribbon
(10 cm x 0.8 mm x 18 IJm) and measuring the voltage across the wire
CTejedor et al. , 2000). The impedance of the metallic glass is found to be
linearly decreasing with applied stress, and the impedance change is about
10% over the range of 500 MPa. Based on the Sl effect, the metallic glasses
are also incorporated in multi-vibrator circuits for the detection of blood vessel
pulsation (Kusomoto et aI., 1999; Mohri et al. , 2001).

11.3.3 Magnetostrictive Delay Line Sensors


Metallic glasses, especially Fe-rich alloys, are widely used to construct
magnetostrictive delay lines (MOL) because they have a large magnetoelastic
coupling, which allows efficient energy conversion between magnetic fields
and acoustic waves. The basic arrangement of the MOL is illustrated in
Fig. 11. 13. The operational principle of the MOL is as follows: A current
pulse is first sent to the excitation coil to generate an ac magnetic field, and an
elastic wave is created and propagated along the MOL due to the
magnetostrictive effect. As the elastic wave propagates to the other end of the
MOL, it generates an induction voltage at the sensing coil due to the inverse
Nano-Structural Magnetoelastic Materials for Sensor Applications 357

magnetostrictive effect. The parameter of interest, such as force, which is


applied between the source and receiver, dissipates the elastic energy of the
propagating acoustic wave, and in turn attenuates the induction voltage
received at the sensing coil. By taking the ratio of the received voltage to the
excitation voltage, the attenuation of the elastic wave can be determined and
be used to measure the applied force. Bias fields are usually applied at the
excitation and receiving regions to enhance the magnetostrictive effect of the
metallic glass.
Delay line
termination Excitation
i1
1
Force
Receiving
Delay line
termination
\[)--tffiID-ttClti_ _-6"_ _ffiIDftiCOttitl ---<<JI

I_N_S~ P
DC bias magnet

Figure 11.13 The basic arrangement of a magnetoelastic delay line (MOL). An elastic
wave is generated by the excitation coil and propagated along the MOL due to the
magnetostrictive effect.

11.3.3. 1 MDL Stress and Torque Sensors


Hristoforou and Reilly (1992) have designed a simple MOL force sensor using
a Metglas 2605SC ribbon. The sensor is similar to that shown in Fig. 11. 13,
but the excitation coil is replaced by a conductor line carrying a voltage pulse.
Experimental results show their MOL force sensor can detect a small force of
O. 01 N, and the force dependency follows an exponential profile up to O. 35 N.
Chiriac et al. (1997a) and Hristoforou et al. (1996) also designed a stress
and torque sensor based on the MOL. The sensor, illustrated in Fig. 11. 14, is
composed of a FeSiB metallic glass wire with its top end fixed and the bottom
end loaded with a mass or torque. An excitation coil is placed near the top end
of the wire to generate an acoustic wave, and a sensing coil is placed at the
bottom end to convert the propagating acoustic wave to an induced voltage.
Two bias coils are used to generate the required biasing field for enhancing the
magnetostrictive effect of the metallic glass. The attenuation of the acoustic
wave, which is due to the mass load or torque, is recorded as the reduction in
the measured induced voltage. Their experimental results show the as-cast
metallic glass has a linear response towards force, and a monotonic but non-
linear response towards torque. They also found annealing the metallic glass at
300C increases the sensitivity but reduces the linearity of the sensor
response.
Recently, Hristoforou et al. (1998) and Chiriac et al. (2001 a) simpl ified
the design of their MOL stress/torque sensor by using overlapping sensing and
excitation coils. Since both ends of the metallic glass wire are bounded, the
acoustic wave generated by the excitation coil will be reflected from the ends
358 Keat G. Ong and Craig A. Grimes

back to the sensing coil. The response of the overlapping coil design is almost
identical to the original design.
11.3.3.2 MDL Displacement and Vibration Sensors
Using MDL setup, metallic glasses are also used to measure the displacement
of an object by attaching the object to a metallic glass ribbon, or attaching the
moving object with a marker magnet. The displacement sensor designed by
Germano et al. (2000), similar to that shown in Fig. 11.14, consists of a long
metallic glass ribbon surrounded by an excitation coil and a receiving coil; the
object of interest is attached to the end of the ribbon. As the object is moving
away from the sensor, it pulls the metallic glass ribbon and changes the
applied stress, altering the attenuation of the acoustic wave and in turn
changing the voltage received by the receiving coil. Germano et al. used
Fe40 Ni40 P14 8 6 , Feso 8 14 Si6 , and Fes2 8 12 Si 4 C2 in their experiment, and they
found Fe4oNi4oP1486 has the best linearity compared to the others.

Detection coil ...fL r----------,

Excitation
circuit

Detection
circuit

Excitation coil

Force

Figure 11.14 An MDL stress and torque sensor.

Another MDL displacement sensor by Germano and Lanotte (1997)


consists of a metallic glass ribbon surrounded by a pair of excitation and
detection coils, but the moving object is a permanent magnet. When the
permanent magnet changes position, the dc magnetic field experienced by the
metallic glass ribbon changes, changing the ribbon susceptibility and the
acoustic wave attenuation.
The displacement sensor developed by Chariac and Marinescu (2000b)
consists of a long metallic glass with an excitation coil at one end (Fig. 11. 15).
The metallic glass is set to resonance so a standing wave is generated along
the sensor. A mobile pick-up coil is then moved along the length of the
sensor. Since the sensor experiences a standing wave vibration, the voltage
Nano-Structural Magnetoelastic Materials for Sensor Applications 359

recorded by the pick-up coil V will be in the form of V (x) = Vo sine 1T x / L)


where x is the displacement. L is the sensor length and V o is a constant
depending on the sensor setup; the displacement x is determ ined from the
measured voltage V.
.. x

Metallic glass

1 ~
r
I

Excitation coil
Biasing coil
~.

Mobile pick-up coil

Figure 11.15 A displacement sensor consisting of a metallic glass set to resonance so a


standing wave is generated along the sensor.

Due to its high sensitivity and fast response. the MDL displacement
sensor is also used to detect small vibrations. Ausanio et al. (2001) have
successfully illustrated the detection of a vibrating steel ribbon by placing an
MDL displacement sensor near the ribbon. The vibration of the steel ribbon
changes the magnetic flux distribution around the metallic glass. and in turn
changes the susceptibility of the metallic glass and the output voltage from the
sensing coil. Ausanio et al. also demonstrated the monitoring of the vibration
of a non-magnetic object by attaching a small magnet on that object.
11. 3. 3. 3 Measurement of Thin-film Thickness
A stress sensor designed for measuring thin film thickness by using the MDL
arrangement was developed by Hristoforou et al. (1999). The sensor. shown
in Fig. 11.16. consists of two metallic glasses at the two ends of a rectangular
glass substrate. An acoustic wave is generated by one of the metallic glasses.
and is propagated along the substrate. When the acoustic wave reaches the
other metallic glass. an induced voltage proportional to the amplitude of the
acoustic wave is generated. Since the acoustic wave attenuation is directly
proportional to the weight on the glass substrate. the thin film thickness can be
determined by measuring the amplitude of the induced voltage.

11.3.4 Magnetoelastic Resonance Sensors


Magnetoelastic resonance sensors operate by monitoring the changes in the
magnetoelastic resonant frequency of a metallic glass ribbon in response to
changing stress and field. Generally. magnetoelastic resonance sensors are
made of Fe-rich metallic glasses because they have a high magnetostriction
coefficient. which allows them to exhibit pronounced mechanical vibrations
360 Keat G. Ong and Craig A. Grimes

Silver paint
Copper ribbon connector Metallic glass film

Pulse Pulse
voltage voltage
generator detector

Sensor support Glass substrate

Figure 11.16 An MOL stress sensor designed for measuring thin film thickness.

when excited by an ac magnetic field, and a high magnetoelastic coupling,


which allows efficient conversion between magnetic energy and elastic
energy.
To determine the resonant frequency of a metallic glass ribbon, an ac
magnetic field is first generated from an excitation coil to vibrate the ribbon.
The frequency response of the sensor in vibration is then recorded as an
induced voltage at the sensing coil (8toyanov and Grimes, 2000). The
excitation and detection of the sensor can be performed in the frequency
domain, where a function generator produces a frequency-varying sinusoidal
wave at the excitation coil, and the voltage across the sensing coil is
measured as a funct ion of frequency (F ig. 11. 17). The sensor can also be
excited and detected in the time domain, where a pulse generator produces a
magnetic pulse, and a digital oscilloscope is used to capture the transient
time-dependent response of the sensor. The time-domain signal is then
converted to the frequency domain using fast Fourier transform, from which the
resonant frequency is determined.
Recently, Jain et al. (2000, 2001a,2001b) proposed the detection of
the resonant frequency of the magnetoelastic resonance sensors acoustically
with a microphone. The advantage of acoustic detection is the longer detection
range; up to 1 m has been reported. In addition to the acoustical detection
method, the sensor can be detected optically by shining a laser beam on the
sensor surface. The vibration of the sensor modulates the intensity of the laser
beam reflected off the vibrating surface, and the changes in intensity are
detected by a phototransistor as a change in voltage that can be monitored
electronically. The advantage of the optical detection method is the large
detection range; up to few meters have been reported. However, the optical
detection method requires precise sensor alignment. The magnetic, acoustic,
and optical detection methods are illustrated in Fig. 11.17.
11. 3. 4. 1 Atmospheric Pressure Measurement
A magnetoelastic resonance sensor designed for the measurement of
atmospheric pressure was developed by Grimes et al. (1999c, 2000a) using
Nano-Structural Magnetoelastic Materials for Sensor Applications 361

Microphone

Low-noise
pre-amp

Function generator Lock-in amplifier


or or
pulse generator digital oscilloscope

Figure 11. 17 The metallic glass sensor can be interrogated magnetically, acoustically,
or optically. Magnetically the sensor is excited by an excitation coil and is detected by a
sensing coil, while acoustically it is detected by a microphone. A laser emitter and a
phototransistor are used to interrogate the sensor optically.

an as-cast Metglas 2826 MB ribbon. The variation of the atmospheric pressure


changes the applied stress on the sensor and in turn changes the resonant
frequency. Experimental results show the Metglas 2826 MB sensor is linearly
dependent upon atmospheric pressure over the range of 0 to 80 psi, with a
resonant frequency shift of 5 Hz/psi. Recently, Kouzoudis and Grimes (2000a)
showed the sensitivity of the pressure sensor could be increased by pre-
stressing the sensor through dimpl ing, curving, or folding the sensor. The
additional stresses on the sensor cause a basal plane vibration in addition to
the longitudinal vibration. Since the surface area of the basal plane is at least
3 orders larger than the cross section of the typical sensor ribbon, the basal
plane vibration significantly increases the sensor interactions with air.
11. 3. 4. 2 Monitoring of Liquid Flow Rate
A metallic glass ribbon can also be used to measure liquid flow rates since the
flowing liquid create a stress, which is proportional to the flow rate, on the
sensor surface and causes changes in the magnetoelastic resonant frequency of
the ribbon. A magnetoelastic resonance sensor for fluid-flow measurement was
fabricated using a Metglas 2826 MB ribbon by Kouzoudis and Grimes (2000b) .
The sensor length is placed parallel to the flow direction of the liquid (Fig. 11. 18a).
The changes in the

Вам также может понравиться