Вы находитесь на странице: 1из 464

MULTISCALE DEFORMATION AND FRACTURE IN MATERIALS

AND STRUCTURES
SOLID MECHANICS AND ITS APPLICATIONS
Volume 84

Series Editor: G.M.L. GLADWELL


Department of Civil Engineering
University of Waterloo
Waterloo, Ontario, Canada N2L 3GI

Aims and Scope of the Series


The fundamental questions arising in mechanics are: Why?, How?, and How much?
The aim of this series is to provide lucid accounts written by authoritative researchers
giving vision and insight in answering these questions on the subject of mechanics as it
relates to solids.

The scope of the series covers the entire spectrum of solid mechanics. Thus it includes
the foundation of mechanics; variational formulations; computational mechanics;
statics, kinematics and dynamics of rigid and elastic bodies: vibrations of solids and
structures; dynamical systems and chaos; the theories of elasticity, plasticity and
viscoelasticity; composite materials; rods, beams, shells and membranes; structural
control and stability; soils, rocks and geomechanics; fracture; tribology; experimental
mechanics; biomechanics and machine design.

The median level of presentation is the first year graduate student. Some texts are mono-
graphs defining the current state of the field; others are accessible to final year under-
graduates; but essentially the emphasis is on readability and clarity.
Multiscale Deformation
and Fracture in Materials
and Structures
The James R. Rice 60th Anniversary Volume

Edited by

T.-J. Chuang
National Institute of Standards & Technology,
Gaithersburg, U.S.A.

and

J. W. Rudnicki
Northwestern University,
Evanston, Illinois, U.S.A.

KLUWER ACADEMIC PUBLISHERS


NEW YORK, BOST ON, DORDRECHT, LONDON, MOSCOW
eBook ISBN 0-306-46952-9
Print ISBN 0-792-36718-9

2002 Kluwer Academic Publishers


New York, Boston, Dordrecht, London, Moscow

All rights reserved

No part of this eBook may be reproduced or transmitted in any form or by any means, electronic,
mechanical, recording, or otherwise, without written consent from the Publisher

Created in the United States of America

Visit Kluwer Online at: http://www.kluweronline.com


and Kluwer's eBookstore at: http://www.ebooks.kluweronline.com
Editors Preface

The work of J. R. Rice has been central to developments in solid mechanics over the last
thirty years. This volume collects 21 articles on deformation and fracture in honor of J.R.
Rice on the occasion of his 60th birthday.Contributors include students (P. M. Anderson,
G. Beltz, T.-J. Chuang, W.J. Drugan, H. Gao, M. Kachanov, V. C. Li, R. M. McMeeking,
S. D. Mesarovic, J. Pan, A. Rubinstein, and J. W. Rudnicki), post-docs (L. B. Sills, Y.
Huang, J.Yu, J.-S. Wang), visiting scholars (B. Cotterell, S. Kubo, H. Riedel) and
co-authors (R. M. Thomson and Z. Suo). These articles provide a window on the diverse
applications of modern solid mechanics to problems of deformation and fracture and insight
into recent developments.
The last thirty years have seen many changes to the practice and applications of
solid mechanics. Some are due to the end of the Cold War and changes in the economy.
The drive for competitiveness has accelerated the need to develop new types of materials
without the costly and time-consuming process of trial and error. An essential element is
a better understanding of the interaction of macroscopic material behavior with microscale
processes, not only mechanical interactions, but also chemical and diffusive mass transfer.
Unprecedented growth in the power of computing has made it possible to attack increasingly
complex problems. In turn, this ability demands more sophisticated and realistic material
models. A consistent theme in modern solid mechanics, and in this volume, is the effort to
integrate information from different size scales. In particular, there is an increasing
emphasis on understanding the role of microstructural and even atomistic processes on
macroscopic material behavior. Despite the great advances in computational power, current
levels do not approach that needed to employ atomic level formulations in practical
applications. Consequently, idealized problems that link behavior at small, even atomic,
size scales to macroscopic behavior remain essential.
It would be presumptuous to hope that the articles here are as original, rigorous,
clear and as strongly connected to observations as the work of the man they are meant to
honor. Nevertheless, we hope that they do reflect the high standards that he has set. That
they do is in no small measure a consequence of the interaction, both formal and informal,
of the authors with J. R. Rice and the inspiration that his work has provided.
The articles in this volume are grouped into sections on Deformation and Fracture
although, obviously, there is some overlap in these topics. As is evident by reading the
titles, the scope and subjects of the articles are diverse. This reflects not only the extensive
impact of Rices work but also the broad applicability of certain fundamental tools of solid
mechanics.
vi EDITORS PREFACE

FRACTURE: Arguably, Rices most well-known contribution is the


introduction of the J-integral in 1968 and its application to problems of fracture. Because
of its path-independent property, the integral has become a standard tool of fracture
mechanics that makes it possible to link processes at the crack-tip to applied loads. Three
of the papers in the Fracture section discuss this J-integral (and several others use it). Kubo
gives a concise catalog of various versions of the integral and related extensions. Li
discusses applications of the J-integral to characterization and tailoring of cementitious
materials. A special feature of these materials is the presence of fibers or aggregate
particles that transmit tractions across the crack-faces behind the tip. In his 1968 paper,
Rice showed that the J-integral is equal to the energy released per unit area of crack advance
for elastic materials. Consequently, this energy or the value of J could be used as criterion
for fracture. Haug and McMeeking use the J-integral to study the effect of an extrinsic
surface charge on the energy release rate for a piezoelectric compact tension specimen.
They find that the presence of the free charge diminishes the effect of the electric field and
suggest that this will complicate attempts to infer the portions of the crack tip singularity
that are due to stress and to the electric field. A related path-independent integral, the
M-integral, is used by Banks-Sills and Boniface to determine the stress intensity factors for
a crack on the interface between two transversely isotropic materials. A finite element
analysis is used to determine the asymptotic near-field displacements needed to evaluate the
M-integral.
Interpretation of the J-integral as an energy release is rigorous only for nonlinear
elastic materials. But much of its usefulness arises from applications to elastic-plastic
materials whose response, for proportional loading paths, is indistinguishable from a
hypothetical nonlinear elastic one. For significant deviations from proportional loading, the
interpretation of J in terms of fracture energy is approximate. Cotterell et al. present a
method for accounting for the extra work arising from deviations from proportional loading
due to significant crack growth in elastic plastic materials.
Crack growth is affected not only by mechanical loading (or coupled piezoelectric
loading as considered by Haug and McMeeking) but also by chemical processes. Numerical
simulations by Tang et al. show that the presence of chemical activity at the crack tip can
lead to blunting, stable steady crack growth or unstable sharpening of the crack tip. In the
steady state regime, the computed crack velocity as a function of applied load agrees
qualitatively with experiments but uncertainties in material parameters make quantitative
comparison difficult. Consistent with previous studies, Tang et al. find the existence of a
threshold stress level that leads to sharpening and fracture, but, contrary to previous studies,
this threshold depends not only on the mechanical driving force, but also on the chemical
kinetics.
A classic problem of material behavior is to delineate the conditions for which
materials fail ductilely or brittlely. Rice and Thomson addressed this problem by
considering the interaction of a dislocation with a sharp crack-tip and arguing that ductile
behavior occurred when the energetics of the interaction favored emission of a dislocation.
In a concise analysis, Beltz and Fischer extend this formulation to consider the effect of the
T-stress, that is , the non-singular portion of the crack-tip stress field. They show that the
EDITORS PREFACE vii

effect of this stress can be significant for small cracks, with lengths on the order of 100
atomic spacings.
Klein and Gao present an innovative approach to the problem of dynamic fracture
instability. They suggest that the discrepancy between predictions and observations could
be resolved by including non-linear deformations near the crack-tip. They do this by a
cohesive potential model that bridges the gap between continuum scale and atomistic scale
calculations. Using as a measure of failure the loss of strong ellipticity, they suggest that
crack branching may be associated with a loss of stiffness in biaxial stretching near the
crack-tip.
Several pioneering papers by Rice have considered the problem of determining the
stress and deformation fields near the tip of a crack in a ductile material. The chapter by
Drugan extends consideration to the case of a crack propagating along the interface of two
ductile (elastic-ideally plastic) materials. An interesting by-product of the analysis for
anti-plane deformation of bimaterials is a family of admissible solutions for homogeneous
materials (including the well-known Chitaley -McClintock solution). Analysis reveals that
beyond a certain level of material mismatch (ratio of yield stresses) a single term of the
asymptotic expansion is not sufficient to characterize accurately the near-tip field. This
suggests that the number of terms required will depend on some microstructural distance.
Yu and Cho present detailed observations of the crack-tip fields in plastically
deforming copper single crystals and compare them with fields predicted by Rice
(Mechanics of Materials, 1987). They suggest that discrepancies could be due to absence
of latent hardening in the elastic ideally plastic model analyzed by Rice.
Rubinstein presents the results of numerical calculations based on a complex
variable formulation for a variety of micromechanical models of composites. Though the
calculations are elastic, they take explicit account of various reinforcing fibers, particles,
etc. and, as a result the solutions depend on the ratio of fiber size to spacing, an important
design variable.
DEFORMATION: Another major contribution of Rice has been the
development of shear localization theory as a model of failure in ductile materials. In
contrast to fracture, where the stress intensification caused by acute geometry plays a
dominant role, the approach of shear localization is based on the constitutive description of
homogeneous deformation. The constitutive relation developed by Gurson, under Rices
direction, has seen much application in this context because it includes softening due to the
nucleation and growth of micro-voids, an important microscale feature of ductile metal
deformation. Chen et al. discuss modifications of the Gurson model that are necessary to
describe the anisotropy of aluminum sheets. A related chapter by Chien et al. uses a three
dimensional finite element analysis of a unit cell to confirm the accuracy of a
phenomenological anisotropic yield condition for porous metal and apply the
phenomenological condition to analyze failure in a fender forming operation. The chapter
by Rudnicki discusses shear localization of porous materials in a quite different context: the
effects of coupling between pore fluid diffusion and deformation on the development of
shear localization in geomaterials.
viii EDITORS PREFACE

Although the constitutive model developed by Gurson and those used by Chien et
al., Chen et al. and Rudnicki are more complex than classic elastic-plastic relations, they
include microstructural information simply by means of the void volume fraction or
porosity. The paper by Riedel and Blug presents an example of the type of sophisticated
constitutive model needed for implementation in a finite element code to model a complex
technology, solid state sintering. Application of the model to silicon carbide demonstrates
the level of detail and accuracy this kind of material modelling combined with finite element
analysis can bring to technological processes.
Elastic-plastic contact is an example of the fruitful application of continuum
mechanics to microscale processes. Applications include indentation hardness testing,
atomic force microscopy, powder compaction, friction and wear. Mesarovic reviews and
summarizes the current understanding in this area and identifies a number of problems in
need of further work. Recent computational advances have improved understanding but
further work is needed in several areas.
Hydrogen is an element whose presence on an interface or at a crack-tip can lead
to embrittlement. In an elegant analysis that combined thermodynamics and fracture
mechanics and extended the introduction of surface energy into fracture analysis by Griffith,
Rice showed how the presence and mobility of segregants can alter the surface energy.Wang
reviews the analysis of Rice and co-workers and shows that the predictions are consistent
with observations of hydrogen embrittlement in iron single crystals.
Anderson and Xin address the classic problem of the stress needed to drive a
dislocation. In particular, they examine how this stress is affected by a welded interface
using a model that allows them to vary independently the unstable stacking fault energy gus,
the peak shear strength and the slip at peak shear. Using a numerical solution, they find that
the critical resolved shear stress increases with gus, but is relatively insensitive to the
maximum shear strength.
Suo and Lu present a model for the growth of a two-phase epilayer on an elastic
substrate. By means of a linear perturbation analysis and numerical computations, they
show that the competition between phase coarsening, due to phase boundary energy, and
phase refining, due to concentration dependent surface stress, can lead to a variety of growth
patterns, including a stable periodic structure.
The chapter by Kachanov et al. gives a complete solution for the problem of
translation and rotation of ellipsoidal inclusions in an elastic space. Although they do not
pursue applications of the solution, the solution is relevant to deformation around hard
particles in a matrix, motion of embedded anchors, etc.
Thomson et al. present a percolation theory approach to addressing the inevitable
inhomogeneous deformation on the microscale. They show how it can be used to construct
stress/ strain response and give insight into processes of microlocalization.
We consider it an honor and privilege to have had the opportunity to edit this
volume. In the preparation of the biography, H. Gao, W. Drugan and Y. Ben-Zion provided
extra needed information. Jim himself provided autobiographical source material and
helped proofread it to assure its correctness and completeness. We are grateful to the
individual authors for their contributions and timely cooperation, and to the technical review
EDITORS PREFACE ix

board members who enhanced the quality of the volume by providing critical reviews on the
articles.
Our special thanks are due to Kluwer Academic Publishers, Dordrecht Office and
its professional staff for their editing and production, and for their agreement to publish the
Volume given even when it was still unwritten, but existed simply as a proposal in the form
of a list of authors and titles. Financial support and encouragement from NIST management
team, S. Freiman, G. White and E. R. Fuller, Jr. are gratefully acknowledged. Finally, we
would like to express our appreciation to Drs. W. Luecke, X. Gu and J. Guyer for their help
in the editing of this book.

T-J. CHUANG, Gaithersburg, MD J. W. RUDNICKI, Evanston, IL


25 August 2000
This page intentionally left blank.
TABLE OF CONTENTS

Editors Preface v

Biography of James R. Rice xv


T.-J. Chuang and J. W. Rudnicki

List of Publications by James R. Rice xxvii

List of Contributors xli

PART I: DEFORMATION

Approximate Yield Criterion for Anisotropic Porous Sheet Metals and its 1
Applications to Failure Prediction of Sheet Metals under Forming Processes
W. Y. Chien, H.-M. Huang, J. Pan and S. C. Tang

A Dilatational Plasticity Theory for Aluminum Sheets 17


B. Chen, P. D. Wu, Z. C. Xia, S. R. MacEwan, S. C. Tang and Y. Huang

Internal Hydrogen-Induced Embrittlement in Iron Single Crystals 31


J.-S. Wang

A Comprehensive Model for Solid State Sintering and its Application 49


to Silicon Carbide
H. Riedel and B. Blug

Mapping the Elastic-Plastic Contact and Adhesion 71


S. Dj. Mesarovic

The Critical Shear Stress to Transmit a Peierls Screw Dislocation 87


across a Non-Slipping Interface
P. M. Anderson and X.J. Xin
xii TABLE OF CONTENTS

Self-Organizing Nanophases on a Solid Surface 107


Z. Suo and W. Lu

Elastic Space Containing a Rigid Ellipsoidal Inclusion 123


Subjected to Translation and Rotation
M. Kachanov, E. Karapetian, and I. Sevostianov

Strain Percolation in Metal Deformation 145


R. M. Thomson, L. E. Levine and Y. Shim

Diffusive Instabilities in Dilating and Compacting Geomaterials 159


J. W. Rudnicki

PART II: FRACTURE

Fracture Mechanics of an Interface Crack between a Special Pair of 183


Transversely Isotropic Materials
L. Banks-Sills and V. Boniface

Path-Independent Integrals Related to the J-Integral and Their Evaluations 205


S. Kubo

On the Extension of the JR Concept to Significant Crack Growth 223


B. Cotterell, Z. Chen and A. G. Atkins

Effect of T-Stress on Edge Dislocation Formation at a Crack Tip under 237


Mode I Loading
G. E. Beltz and L. L. Fischer

Elastic-Plastic Crack Growth along Ductile/Ductile Interfaces 243


W. J. Drugan

Study of Crack Dynamics Using Virtual Internal Bond Method 275


P. A. Klein and H. Gao

Crack Tip Plasticity in Copper Single Crystals 311


J. Yu and J. W. Cho
TABLE OF CONTENTS xiii

Numerical Simulations of SubCritical Crack Growth 331


by Stress Corrosion in an Elastic Solid
Z. Tang, A. F. Bower and T.-J. Chuang

Energy Release Rate for a Crack with Extrinsic Surface Charge 349
in a Piezoelectric Compact Tension Specimen
A. Haug and R. M. McMeeking

Micromechanics of Failure in Composites 361


-An Analytical Study
A. A. Rubinstein

J-Integral Applications to Characterization and Tailoring 385


of Cementitious Materials
V. C. Li

Author Index 407

Subject Index 415


James R. Rice
Biography of James R. Rice

James Robert Rice (JRR) was born on 3 December 1940 in Frederick, Maryland to Donald
Blessing Rice and Mary Celia (Santangelo) Rice. Located some 50 miles northwest of the
nations capital, Frederick was then a small city of about 20,000 people, set in a rural,
farming area. Commemorated in Whittiers poem about Dame Barbara Fritchies patriotism,
Frederick was a crossroads for troop movements during the Civil War (1861-1865) and the
birthplace of Francis Scott Key who wrote the American National Anthem. JRRs mother
Mary was the child of a Sicilian immigrant family and now resides in Adamstown,
Maryland. The family of JRRs father, Donald, had long lived in that part of the USA.
Donald, who died in 1987, operated a gasoline station, served 3 terms as alderman and a
term as mayor of Frederick City in the early 1950s, later founded a successful tire company,
and, like Mary, was highly active in Frederick community affairs.
JRR was raised in Frederick, and was the second of three children. His older
brother, Donald Blessing Rice Jr., served as corporate CEO of several companies (such as
the RAND Corporation) in the private sector and one term as Secretary of the U.S. Air
Force under the Bush Administration. He now resides in Los Angeles. JRRs younger
brother, Kenneth Walter Rice, continues to live in Frederick and runs the business started
by his father.
JRR attended primary and secondary school at St. Johns Literary Institute, a local
parish school in Frederick. He played baseball and basketball, worked part-time delivering
newspapers and in his fathers businesses, and read a lot. Influenced by his high school
teachers of math and physics, recruited from Fort Dieterich, a local army base, JRRs early
interest in auto mechanics gradually evolved into an interest in mechanical engineering.
Armed with several scholarships, he began undergraduate studies in that subject at Lehigh
University in Bethlehem, PA, in 1958, one year after the launch of Sputnik propelled the
U.S. into a keen competition in outer space with the then-USSR.
During his undergraduate studies at Lehigh, JRR realized his particular interest was
in theoretical mechanics, especially fluid and solid mechanics, and applied mathematics.
Under the influence of inspiring teachers including Ferdinand Beer, Fazil Erdogan, Paul
Paris, Jerzy Owczarek, George Sih, and Gerry Smith, he did his subsequent studies in the
engineering mechanics and applied mechanics programs. Paul Paris has said that for the
courses JRR took from him, half of Pauls preparation for each lecture consisted of
answering the questions JRR had posed during the previous class meeting. Because of his
proficiency in math and physics, JRR earned all his academic degrees, from B.S. to Ph.D
in only six years (1958-1964), the shortest time in Lehighs record. Ferdinand Beer directed
JRRs M.S. and Ph.D. theses on stochastic processes, specifically on the statistics of highly
correlated noise. The results were summarized in 1964 in his Ph.D. thesis, entitled
Theoretical Prediction of Some Statistical Characteristics of Random Loadings Relevant
xvi BIOGRAPHY OF J. R. RICE
to Fatigue and Fracture. At the same time, he continued working with George Sih on the
subject of his undergraduate research project, elastic stress analysis of cracks along a bi-
material interface. He independently developed a simple elastic-plastic crack model, which
turned out to be the same as D. S. Dugdale had already published, and then extended the
model to the case of cyclic loads. His work on The Mechanics of Crack Tip Deformation
and Extension by Fatigue was published in ASTM STP 415 in 1967, and was awarded the
ASTM Charles B. Dudley Medal in 1969.
In the late 1950s, fracture mechanics was still in the early stages of development.
Egon Orowan of MIT and George Irwin of Naval Research Laboratory were beginning to
advocate using stress analysis of cracks to solve fracture and fatigue problems in
conventional metals and metal alloys. Motivated by the problems encountered while
working at Boeing in the summers, Paul Paris was especially keen to work in this field.
Together Paris, George Sih and Erdogan offered the first graduate course on fracture
mechanics, which JRR took in his senior year. In addition, they recruited bright graduate
students, including JRR, to do thesis research in this area. This environment cultivated
JRRs interest in fracture mechanics, which became a major focus of his teaching and
research.
After JRRs graduation from Lehigh in 1964, his advisor, Ferdinand Beer,
suggested he accept an offer from Daniel C. Drucker to be a post-doctoral research fellow
in the Solid Mechanics Group of the Division of Engineering at Brown University. Brown
was (and still is) well known internationally in the solid mechanics community. At that time
many world-renowned researchers in solid mechanics were members of the faculty. They
included, among others, Daniel C. Drucker, Morton E. Gurtin, Harry Kolsky, Joseph
Kestin, Alan C. Pipkin, Ronald S. Rivlin, Richard T. Shield, and Paul S. Symonds.
At Brown, JRR, armed with enthusiasm, energy, and innovative ideas, pursued his
research on many critical fronts in fracture mechanics. He continued to collaborate with his
former professors on the unfinished work from Lehigh, including characterization of fatigue
loadings, plastic yielding at a crack tip and stress analysis of cracks and notches in elastic
and work-hardening plastic materials under longitudinal shear loading. At Lehigh, he had
also obtained some results for determining energy changes due to material removal, such
as cracking or cavitation, in a linear elastic solid. At Brown, Drucker opened his eyes to the
importance of generalizing these results to the widest possible class of materials; thus, JRR
developed this work into a procedure for calculating energy changes in a general class of
solids. This work led to JRRs discovery of the well-known J-integral a few years later.
With these impressive achievements, he was offered a tenure-track faculty job as Assistant
Professor in 1965.
As an assistant professor at Brown, JRR devoted his energy and efforts not only
to research but also to teaching. He always believed that a good professor must excel in
teaching and research. He offered many courses in applied mechanics. He developed his
own lecture notes in each course without relying on specific text books. During lecturing
in a typical class, he memorized every important piece of information and used the
blackboard to convey the concepts to students. He was an excellent and effective
BIOGRAPHY OF J. R. RICE xvii

communicator. Students were always welcome and encouraged to ask questions or engage
in discussions. Copies of his lecture notes highlighting the key information including
methods of derivations and final resulting formulae were distributed to his students.
In research, he obtained federal funding from agencies such as NSF, DARPA,
NASA, ONR, and the DOE to support project initiatives on mechanics of deformation and
fracture. At this time, fracture mechanics was still in the early stages of development. JRR
seized the opportunity to work out many unsolved problems in stress and deformation fields
around a crack in various materials systems, mostly in 2D. Some examples are: elastic-
plastic mechanics of crack extension, stresses in an infinite strip containing a semi-infinite
crack, plane-strain deformation near a crack in a power-law hardening material (with G.F.
Rosengren), energy changes in stressed bodies due to void and crack growth (with D.C.
Drucker), a path independent integral and the approximate analysis of strain concentration
by notches and cracks. At the invitation of H. Liebowitz, this work was summarized in a
classic review article entitled Mathematical Analysis in the Mechanics of Fracture, which
appeared in 1968 as Chapter 3, in Volume 2, Mathematical Fundamentals of Fracture, of
the book series, Fracture: An Advanced Treatise.
Of particular significance was the discovery of a path-independent integral
resulting from his prior probe into energy variations due to cracking of a nonlinear elastic
solid. He named this particular integral the J-Integral with the upper case letter J
inadvertently coinciding with his nickname big Jim respectfully used by his students. This
integral turned out to coincide with a 2D version of the general 3D energy momentum tensor
proposed by J. D. Eshelby in England in 1956. A similar concept was also developed by
Cherepanov in Russia at about the same time as Rices J-integral, but JRR exploited the
integrals usefulness more fully in fracture analysis, especially by focusing on aspects
relating to path-independence. Because of its path independence, the J-integral is a powerful
tool to evaluate energy release due to cracking, bypassing the difficulties arising from strain
concentration at the crack-tip. Using the procedure he developed with Drucker, JRR showed
that the J-Integral is identical to the rate of reduction of potential energy with respect to
crack extension. In addition, JRR, together with the late Gran F. Rosengren, showed in
1968 that the J-integral plays the role of a single unique parameter that governs the
amplitude of the nonlinear deformation and stress fields inside the plastic zone near a crack
tip. This result established criticality of the J-integral as a criterion for fracture even for an
elastic-plastic material and made possible its use for practical engineering applications.
Simultaneously, John Hutchinson at Harvard also derived a similar result. Based on their
studies, the nonlinear stress distribution in the crack tip zone is now referred to as the
Huchinson-Rice-Rosengren or HRR field. Over the next decade, criticality of the J-
Integral was adopted as the major design criterion against failure. It is used in the ASME
Pressure Vessel and Piping Design Code, and in general purpose finite element codes such
as ABAQUS and ANSYS. JRRs paper on the J-integral, which appeared in the Journal of
Applied Mechanics in 1968, received the ASME Henry Hess Award in 1969 and has
become a classic, attracting more than 1000 citations and references. The J-Integral forms
an essential part of the subject matter contained in any textbook on fracture mechanics.
xviii BIOGRAPHY OF J. R. RICE

Because of this and other contributions, JRR was promoted to Associate Professor in
Engineering in 1968 and received the ASME Pi Tau Sigma Gold Medal Award for
outstanding achievement in mechanical engineering within 10 years following graduation
in 1971.
As Associate Professor at Brown, JRR extended his research interests from
mechanics to the physics and thermodynamics aspects of fracture phenomena. He worked
with his student N. Levy on the prediction of temperature rise by plastic deformation at a
moving or stationary crack-tip. When applied to a set of aluminum and mild steel alloys, this
work helped to explain the experimentally observed relationship between the temperature-
dependent toughness and the loading rate. Other accomplishments included his work with
his student Dennis Tracey on the ductile void growth in a triaxial stress field. This work
clarified the mechanism of void growth under applied stress in ductile metals. The role of
large crack tip geometry changes in plane strain fracture was quantified in a paper with M.
Johnson. He also actively participated in the development of formulations for finite element
computations. He directed Ph.D. thesis research in computational fracture mechanics by
Dennis Tracey. He interacted with Pedro Marcal, a faculty colleague and the founding
developer of the MARC finite element code, and with Dave Hibbitt, Marcals graduate
student and the co-developer of the ABAQUS code. Together, they developed an
appropriate numerical algorithm to compute large strains and large displacements in the
finite element code. This scheme has been implemented in many general purpose finite
element codes such as MARC, ABAQUS and ANSYS.
With another faculty colleague, Joseph Kestin, JRR worked on the application of
thermodynamics to strained solids. For example, although the chemical potential is well-
defined in fluids, the proper definition in solids is not clear. A paper by Kestin and Rice
helped to clarify the concept and served as a starting point to extend JRRs developing
interest in high temperature fracture, namely, creep and creep rupture.
In 1970, JRR was promoted to Full Professor of Engineering. With financial
support from federal funding agencies such as the National Aeronautic and Space
Administration (NASA), Office of Naval Research (ONR), DARPA, National Science
Foundation (NSF) and Atomic Energy Commission (AEC, the predecessor of ERDA and
the Department of Energy (DOE)), he was directing a research team of 7 Ph.D. graduate
students. The team participated in the Materials Research Laboratory, a large-scale,
interdisciplinary research program, funded by DARPA and NSF, and in a program of the
AEC Basic Sciences Division directed by Joseph Gurland. JRRs students worked in a
wide range of areas in the mechanics of solids and fracture: Dennis Tracey, Dave Parks, and
Bob McMeeking in (1) theoretical and computational fracture mechanics; Art Gurson in (2)
constitutive relationships in metals and metallic alloys; Glenn Brown and (Jerry) T.- j.
Chuang in (3) creep and creep rupture in the high temperature range; and Mike Cleary in
(4) mechanics of geomaterials. Representative work in (1) included an alternative
formulation of Bueckners (1970) weight function method to evaluate the stress intensity
factor K I of a given 2D linear elastic cracked solid subject to arbitrary loading, based on any
known solution to the same geometry; a finite element analysis of small scale yielding near
BIOGRAPHY OF J. R. RICE xix

a crack in plane-strain (with N. Levy, P.V. Marcal and W.J.Ostergren); an approximate


method for analysis of a part-through surface crack in an elastic plate (with N. Levy); and
3D elastic-plastic stress analysis for fracture mechanics (with N. Levy and P. V. Marcal).
In (2) JRR worked out the fundamental structure for the time-dependent stress-strain
relationship of a metal in the plastic deformation range and proposed an internal variable
theory for the inelastic constitutive relations in metal plasticity.
In 1971-72, JRR took a year of sabbatical leave with support from a NSF Senior
Postdoctoral Fellowship. He spent the year at the Department of Applied Mathematics and
Theoretical Physics of the University of Cambridge, where he was affiliated with Churchill
College under the support of a Churchill College Overseas Fellowship. At Cambridge, he
worked with a number of people, including Rodney Hill, one of the pioneers in classical
plasticity, Andrew C. Palmer in soil mechanics, and John Knott and his student Rob Ritchie
on elastic-plastic fracture. With Hill, JRR developed a general structure of inelastic
constitutive relations assuming the existence of elastic potentials, and gave a special
implementation for elastic/plastic crystals at finite strain. In the latter case, crystallographic
slip along a set of active slip planes was considered as the sole deformation mechanism
responsible for the inelastic behavior. This theory successfully explained various aspects of
plasticity such as strain hardening, the existence of a flow rule and normality. With Knott
and Ritchie, JRR proposed a relationship between the critical tensile stress and the fracture
toughness of mild steel. The analysis predicts the observed temperature dependence of K IC
in the brittle to ductile transition range. With Andrew Palmer, JRR used his newly
developed J-integral to develop a mode-II shear crack model for the growth of slip
surfaces in over-consolidated clay slopes.
Returning to Brown in 1972, JRR continued to pursue research on many aspects
of fracture mechanics. John Landes and Jim Begley of the Westinghouse R&D Center
became keen advocates of using the J-Integral as a design criterion in the nuclear energy
business, and in a paper with Landes and Paul Paris, JRR developed an elegantly simple
procedure to estimate the value of J-Integrals from experiments. Eventually, this procedure
became the ASTM standard and part of the ASME Pressure Vessels and Piping design code.
Besides analysis on the continuum level, JRR strongly felt that there was a need to study
fracture at the microstructural level in order to bridge the atomic and engineering scales.
One important area that required such a treatment is high temperature creep and creep
rupture where mass transport plays an important role. At that time, a group at Harvard led
by Mike Ashby was also interested in this topic. As a result, there was much interaction
between Harvard and Brown during 1972-74: JRR and Ashby and their students made
frequent mutual visits to give seminars and to exchange ideas. One important result, jointly
developed in 1973 with his student, T.- J. Chuang, was the discovery of creep crack-like
cavity shapes induced by surface diffusion. This type of cavity, referred to as a Chuang-Rice
crack-like cavity, is frequently observed at the grain boundaries of a ruptured tensile
specimen. This work defines the boundary conditions at the cavity apex and satisfactorily
explained non-linear stress dependence on cavity growth rate. The degree of non-linearity
depends on the deformability of the grains, and JRR obtained solutions for the stress
xx BIOGRAPHY OF J. R. RICE

dependence on creep cavity growth in rigid grains (with Chuang, Kagawa, Sills and Sham),
in elastic grains (with Chuang) and in plastic grains (with Needleman). The predicted stress
dependence was verified experimentally by Bill Nix and his students at Stanford in the late
1970s using implanted water vapor cavities at grain boundaries in pure silver and nickel-tin
alloys. Later in the 1980s and 90s, this work was used by many researchers to predict cavity
growth induced by electromigration in aluminum interconnect wires.
In 1973, JRR was offered a Chair by the Brown President, Donald Hornig with the
title L. Herbert Ballou Professor of Theoretical and Applied Mechanics. This privileged title
is an honor comparable to a University Professorship, which is the highest rank of teaching
professors at Brown.
In physical metallurgy, it had become well-known that dislocations at the atomic
level are fully responsible for the room temperature plastic behavior in metals. Since the
early 1960s, many researchers (such as Hirth, Lothe, Mura and Weertman) devoted their
efforts to this area and helped to build the foundation of dislocation theory. JRR was among
those cutting edge scholars who excelled in mathematical dislocation theory. In 1972, he
met Robb Thomson of SUNY-Stony Brook at a conference and they puzzled over the
ductile versus brittle transition phenomenon in crystals. Since dislocation movement leads
to ductility and rapid crack growth leads to catastrophic failure, they believed the
interactions of both must play a dominant role in ductile/brittle behavior. They proposed
that the ability to emit dislocations from a pre-existing sharp crack tip is the source of
ductility in metals. On the other hand, the resistance of a crack tip to dislocation emission
leads to brittleness in ionic or covalent crystals like ceramics. By analyzing the energetic
forces between a dislocation and a crack, they derived an important parameter that governs
the ductility. If this parameter, which is shear modulus times Burgers vector over surface
energy, exceeds 8.5 to 10, then the crystal exhibits intrinsically brittle behavior. If less, it
is generally ductile. The Rice -Thomson theory has become a classic in the Science Citation
Index with more than 200 citations. In the late 1970s, Mike Ohr of Oak Ridge National
Laboratory provided direct experimental evidence for the theory by observing emission of
dislocations from the crack tip in a variety of metal specimens in situ under TEM.
In another noteworthy work, JRR helped his student Art Gurson to develop in 1975
the plasticity theory of porous media, in which yield criteria and flow rules were predicted
in stress space using 2D or 3D unit cell models. The model predicts the effect of porosity
on the plastic behavior of ductile materials and has come to be known as the Gurson
model. It is well-known in the metallurgy and mechanics communities and is one of the
major yield criteria adopted in the commercial general purpose finite element codes for
assessing inelastic behavior of metallic materials.
Motivated by his studies of shear bands with Andrew Palmer, JRR became
interested in the fundamental question of why deformation would localize in a narrow zone.
A basic premise of fracture mechanics, going back to the ideas of Griffith, is that the
presence of flaws in a material causes a local elevation of the stress and leads to propagation
of the flaw and, eventually, to failure. Although this process provides a satisfactory
explanation of failure in many materials, it does not explain why macroscopically uniform
BIOGRAPHY OF J. R. RICE xxi

deformation should give way to localized deformation in very ductile materials or under
conditions of compressive stress that suppress flaw propagation. Based on antecedents in
the work of Hadamard, Hill, Thomas and Mandel, JRR and his student Rudnicki treated the
initiation of localized deformation as a bifurcation from homogeneous deformation and
showed that its onset was promoted by certain subtle features of the constitutive behavior.
This work, which was published in the Journal of the Mechanics and Physics of Solids in
1975, received the Award for Outstanding Research in Rock Mechanics from the U. S.
National Committee on Rock Mechanics in 1977. Although this work was originally
intended to describe fault formation in rock, JRR extended the approach to consider
localized necking in thin sheets (with S. Storen), strain localization in ductile single crystals
(with R. J. Asaro), and limits to ductility in sheet metal forming (with A. Needleman). He
summarized the state of the subject in a keynote lecture on The Localization of Plastic
Deformation at the 14 th International Congress on Theoretical and Applied Mechanics in
Delft in 1976. The printed version of this lecture is a widely-cited classic.
In the early 1970s, there were many reports of observations precursory to
earthquakes that were attributed to the coupling of deformation with the diffusion of pore
fluid. A series of papers, by JRR with students (Cleary and Rudnicki) and Don Simons, an
Assistant Professor at Brown, analyzed the effects of this coupling on models for earthquake
instability and for quasi-statically propagating creep events. One of these papers (Some
basic stress-diffusion solutions for fluid-saturated elastic porous media with compressible
constituents, with M. P. Cleary, Rev. Geophys. Space Phys., 14, pp. 227-241, 1976)
reformulated, in a particularly insightful way, the equations first derived by Biot for a linear
elastic, porous, fluid-infiltrated solid. This version of the equations has proven so
advantageous that it is now the standard form. The models of the earthquake instability
formulated to study these effects were among the first in which the instability was not
postulated but arose in a mechanically consistent way from the interaction of the fault zone
material behavior and the surroundings.
JRRs interest in the mechanics of earthquakes proved durable and became a major
branch of his work. With Florian Lehner and Victor Li, he worked on time-dependent
effects due to coupling of the shallow, elastic portion of the Earths lithosphere with deeper
viscoelastic portions. This work was based on a generalization of an earlier thin plate model
by Elsasser. This work demonstrated that the viscous deformation of the lower crust and
upper mantle following large earthquakes could affect surface deformation for decades and
provided a new model for the interpretation of increasingly detailed surface deformation
measurements. In the early 90s, JRR used the finite element code ABAQUS together with
Yehuda Ben-Zion, Renata Dmowska, Mark Linker, and Mark Taylor to explore the
behavior of this model in 3D and to compare model predictions with geophysical
observations. JRRs growing interest in the mechanics of earthquakes complemented nicely
the interests of his spouse, Renata Dmowska, a seismologist. Together, Renata's analysis
of data and JRRs mathematical models have been combined in several papers on aspects
of earthquakes, particularly in subduction zones.
JRRs interest in the mechanics of earthquakes soon led to a study of frictional
xxii BIOGRAPHY OF J. R. RICE

stability. Stick-slip is a widely observed phenomenon and has long been regarded as a
physical analog for the earthquake instability. But the standard constitutive description,
static and dynamic friction, was inconsistent with the steady sliding often observed and
contained no mechanism for restrengthening that would allow repeated events on the same
surface. Based on experimental observations of Dieterich at the U.S. Geological Survey,
JRR and his student Andy Ruina formulated a rate- and state-dependent constitutive
relationship for sliding on a frictional surface. By examining the stability of a one degree-of-
freedom system with this relationship, they were able to predict the variety of behaviors
observed in rock friction experiments: steady sliding, damped oscillations, stick-slip and
sustained periodic oscillations. Other papers with Tse and Gu examined the dynamics and
nonlinear stability of these systems. JRR and his student Tse showed that when this type of
relationship was applied on a surface between two elastic solids and modified to include a
depth dependence appropriate for the temperature and pressure dependence in the earth, the
calculations produced periodic events with a depth dependence remarkably similar to that
of observed earthquakes.
In the late 1970s and early 1980s, JRR also continued to work on many aspects of
inelasticity and fracture. With Joop Nagtegaal and Dave Parks, he developed a numerical
scheme to improve the accuracy of finite element computations in the fully plastic range.
With Bob McMeeking, he worked out the proper finite element formulation in the large
elastic-plastic deformation regime. With a colleague at Brown, Ben Freund, and a student,
Dave Parks, he helped solve the problem of a running crack in a pressurized pipeline. In
materials science, he studied stress corrosion and hydrogen embrittlement problems.
He also orchestrated a remarkable multidirectional attack on the problem of quasi-
static crack growth in elastic-plastic materials. This began with a paper with Paul Sorensen
in 1978 that proposed an elegant way of using near-tip elastic-plastic fields to derive
theoretical predictions for crack growth resistance curves (J R curves). Then, he and his
student Walt Drugan derived asymptotic analytical elastic-ideally plastic solutions for the
stress and deformation fields near a plane strain growing crack which showed the necessity
of an elastic unloading sector in the near-tip field. [Independent work by L. I. Slepyan in
the then-USSR and Y. C. Gao in China also addressed this problem, for incompressible
material and steady-state conditions.] The detailed numerical finite element elastic-plastic
growing crack solutions of JRRs student T-L. Sham confirmed the analytical predictions,
and in a 1980 paper with Drugan and Sham, JRR combined the method proposed earlier
with Sorensen, with the new analytical asymptotic solutions and Shams numerical results,
to produce a comprehensive and fundamentals-based model of stable ductile crack growth
and predictions of plane strain crack growth resistance curves. Then, with Lawrence
Hermann, JRR conducted and analyzed plane strain crack growth tests and showed that
this theory was indeed capable of describing the experimentally-measured crack growth
resistance curves under contained yielding conditions.
The asymptotic analysis of elastic-ideally plastic growing crack fields, involving
the assembling of different possible types of near-tip solution sectors into complete near-tip
solutions, prompted JRR and Drugan to inquire more fundamentally about what continuity
BIOGRAPHY OF J. R. RICE xxiii

and jump conditions are required across quasi-statically propagating surfaces in elastic-
plastic materials by the fundamental laws of continuum mechanics and broad, realistic
constitutive constraints (such as the maximum plastic work inequality). Their resulting
restrictions (published in the D. C. Drucker Anniversary Volume), and the later
generalization of these to dynamic conditions by Drugan and Shen, have been utilized
repeatedly in elastic-plastic crack growth studies. Not surprisingly, perhaps the most
important applications of these discontinuity results are due to JRR himself, in his
fundamental studies of stationary and growing crack fields in ductile single crystals, wherein
JRR showed that a precise understanding of possible discontinuity types is absolutely
essential in deriving correct solutions. Beginning in 1985 with his student R. Nikolic on the
anti-plane shear crack problem, and in a landmark, pioneering 1987 paper on plane strain
tensile cracks, JRR produced fascinating analytical solutions for the near-tip fields in
elastic-ideally plastic ductile single crystals. These fields differ dramatically from crack
fields in isotropic (i.e., polycrystalline) ductile materials, being characterized by
discontinuous displacements and stresses for stationary cracks, discontinuous velocities for
quasi-statically growing cracks, and, in another fascinating paper with Nikolic in 1988, JRR
showed that the near-tip field for a dynamically propagating anti-plane shear crack in a
ductile single crystal must involve shock surfaces across which stress and velocity jump.
JRR and his student M. Saeedvafa generalized the stationary crack ductile single crystal
solutions to incorporate Taylor hardening, revealing even more complex near-tip behavior.
Other major work in the late 1970s and early 1980s included two important papers
with visiting faculty members: one on the crack tip stress and deformation fields for a crack
in a creeping solid, with Hermann Riedel; and another heavily-cited paper on crack curving
and kinking in elastic materials, with Brian Cotterell.
For his significant contributions to sciences and engineering, JRR was elected to
Fellow grade of the American Academy of Arts and Sciences in 1978, Fellow of the
American Society of Mechanical Engineers and Membership in the National Academy of
Engineering in 1980, and membership in the National Academy of Sciences in 1981.
The next move was to Harvard University in September 1981. A Gordon McKay
Chaired Professorship in Engineering Sciences and Geophysics was created for JRR, jointly
in the Division of Applied Sciences and the Department of Earth and Planetary Sciences.
He further expanded the scope of his research activities along two major branches in
mechanics, namely, fracture of engineering materials and geological materials. At Harvard,
he recruited many bright students from all over the world to work on topical fracture
problems in engineering and geology. He directed Peter Anderson to study constrained
creep cavitation and the Rice-Thomson model, supervised Huajian Gao on three
dimensional crack problems, worked with Jwo Pan, Ruzica Nikolic and Maryam Saeedvafa
on inelastic behaviors of cracks in single crystal metals, and collaborated with Renata
Dmowska, Victor Li, Paul Segall, Andy Ruina, Yehuda Ben-Zion, G. Perrin, J.-c. Gu, Mark
Linker, Simon T. Tse , G. Zheng, and F. K. Lehner in developing friction laws and shear
crack models of geological faults as related to earthquake events in seismology.
JRRs recent work on earthquakes has focused on several important aspects of the
xxiv BIOGRAPHY OF J. R. RICE

process. One issue is the origin of earthquake complexity, that is, the distribution of events
of various sizes, as described by the well-established Gutenberg-Richter relationship. One
previous explanation was that fault slip, as modeled by friction between two elastic solids,
was an inherently chaotic process. In a series of papers that combine elegant analysis and
prodigious calculations, JRR and his post-doc, Yehuda Ben-Zion, showed that the chaotic
behavior predicted in these models was the subtle result of numerical discretization and
oversimplification of the frictional constitutive relation. Other work was motivated by
observations that slip during an earthquake does not propagate in the fashion predicted by
classical dynamic fracture mechanics with most of the surface slipping for the entire
duration of the event. Instead, slip is pulse-like and any point on the surface slips only for
a short time. Papers with Zheng and Perrin showed that only certain types of frictional
constitutive relations were consistent with these observations. Another, very influential
paper, Fault Stress States, Pore pressure Distributions and the Weakness of the San
Andreas Fault addresses a long-standing paradox in earthquake mechanics: A variety of
measurements indicate that the San Andreas fault in southern California is much weaker,
both in an absolute sense and relative to the surrounding crust, than would be expected from
a straightforward interpretation of laboratory friction experiments. JRR showed that the
discrepancy could be resolved by high fluid pressures within the fault zone and summarized
a variety of evidence for this possibility. Another mechanism that can explain the
discrepancy and produce slip in a pulse-like form is dynamic rupture along a bi-material
interface. JRR has been studying this problem recently together with his student K. Ranjith
and Post-Doc A. Cochard, following earlier works of Weertman, Adams, and Andrews and
Ben-Zion, thus returning to a subject he investigated statically as an undergrad at Lehigh.
In the mid-1980s, JRR and other faculty members including John Hutchinson and
Bernie Budiansky formed a joint research team with Tony Evans at the University of
California at Santa Barbara to study mechanical behavior and toughening mechanisms of
ceramics. Between 1988 and 1994, faculty and students at Harvard regularly visited and
exchanged ideas with Tony Evans and his research group at UCSB. The Harvard-UCSB
collaboration generated tremendous research output. During this period, JRR worked with
John Hutchinson, Jian-Sheng Wang, Mark E. Mear and Zhigang Suo on crack growth on
or near a bi-material interface. With Jian-Sheng Wang, he developed a model of interfacial
embrittlement by hydrogen and solute segregation. This model has been referred to as the
Rice-Wang Model which provided a basis for the materials community in pursuit of better
design of steels. Between 1989 and 1995, JRR worked with Glenn Beltz, Y. Sun and L.
Truskinovsky to reformulate the Rice-Thomson model in terms of interactions between a
crack and a Peierls dislocation being emitted from the crack tip. This study eliminated the
need to define a core cut-off radius for dislocations and instead established unstable stacking
fault energy as the new physical parameter governing the intrinsic ductility of crystals.
Rices new model caused an instant sensation among materials scientists and physicists and
is now used as the new paradigm for understanding brittle-ductile transition of crystals.
Separate from his other activities at Harvard, JRR began to develop a growing
interest in three dimensional crack problems, starting around 1984. Together with Huajian
BIOGRAPHY OF J. R. RICE xxv

Gao, the first of his graduate students at Harvard to work on 3-D crack problems, he
developed a series of ingenious methods of analysis based on the idea of 3-D weight
functions, generalizing a 2-D concept he and Hans Bueckner had developed in the early
1970s. These methods were used to study configurational stability of crack fronts, crack
interaction with dislocation loops and transformation strains, and trapping of crack fronts
by tough particles. In 1987, he began to work with K. S. Kim, who spent a year of sabbatical
at Harvard, to generalize these methods to model dynamically propagating 3-D crack fronts.
This then led to a burst of his interests in the following years in the spontaneous dynamics
of 3-D tensile crack propagation and of slip ruptures in earthquake dynamics. He directed
a number of graduate students, post-docs and visiting scientists on those areas, including K.
S. Kim, Yehuda Ben-Zion, G. Perrin, G. Zheng, Phillipe Geubelle, A. Cochard, J. W.
Morrissey, and Nadia Lapusta. He also encouraged other leading scientists such as John
Willis and Daniel Fisher to work in this field. An example of significant discoveries coming
out of these activities is a new kind of wave which propagates along the crack front at a
velocity different from the usual body and surface elastic wave speeds. JRR continues today
to lead an international research effort in crack and fault dynamics. Needless to say, the
output of his research group is of the highest quality and generates significant impact on the
engineering, materials science and geophysics communities.
As a result of his contributions to science and engineering, JRR received numerous
awards and recognitions by professional societies and academic institutions. In 1981, he was
elected to Fellow of AAAS. Next year in 1982, he received the George R. Irwin Medal
from ASTM Committee E-24, shared with John Hutchinson, for significant contributions
to the development of nonlinear fracture mechanics. In 1985, he was one of the recipients
of an Honorary Doctor of Science Degree at his alma mater, Lehigh University. In 1988, he
was elected Fellow of the American Geophysical Union, and received the William Prager
Medal from the Society of Engineering Science for his outstanding achievements in solid
mechanics. Two years later, he was elected Fellow of the American Academy of
Mechanics and the Royal Society of Edinburgh. In 1992, he received an award from AAM
for Distinguished Service to the Field of Theoretical and Applied Mechanics. The
following year he served as Francis Birch Lecturer on Problems on Earthquake Source
Mechanics at the American Geophysics Union. The next year he received the ASME
Timoshenko Medal with the following citation: for seminal contributions to the
understanding of plasticity and fracture of engineering materials and applications in the
development in the computational and experimental methods of broad significance in
mechanical engineering practice. In 1996, he was elected as a Foreign Member of the
Royal Society of London for his work on earthquakes and solid mechanics and received
an honorary degree from Northwestern University. In addition, he received the ASME
Nadai Award for major contributions to the fundamental understanding of plastic flow and
fracture processes in engineering and geophysical materials and for the invention of the J-
Integral which forms the basis for the practical application of nonlinear fracture mechanics
to the development of standards for the safety of structures. He also received the Francis J.
Clamer Medal from the Franklin Institute for Advances in Metallurgy with the citation: for
xxvi BIOGRAPHY OF J. R. RICE

development of the J-Integral for the accurate prediction of elastic-plastic fracture behavior
in metal from easily obtained data. In 1997, he received an honorary Doctor of Science
degree from Brown University. In 1998, a donation from David Hibbitt and Paul Sorensen
of HKS, Inc. established the Rice Professorship at Brown in his honor. Recently, he was
awarded the Blaise Pascal Professorship by the Region Ile-de-France for the 1999 calendar
year for research on Rupture Dynamics in Seismology and Materials Physics, and he was
the recipient of an Honorary Doctoral Degree at the University of Paris VI in March 1999.
He was elected a Foreign Member (Associ trager) of the French Academy of Sciences
in April 2000.
There is no need to place complimentary words here on the impact of his work.
The recognitions described in the previous paragraph speak for themselves. His standards
of scholarship and intellectual honesty are the highest. He is always ready to appreciate the
good work of other colleagues, and to give them proper credit. On the other hand, he does
not hesitate to dispense candid criticism of inconsistent or misguided thinking, though in a
gentle rather than harsh manner -- as some oral comments in conferences or written book
reviews testify.
A man is as young as he thinks. JRR enjoys long walks, whether in urban or
mountain settings, reads broadly in science, history and social commentary, and likes
listening to classical and folk music in his spare time.. He has an excellent sense of humor,
a razor-sharp wit and a cheerful disposition. His wife Renata Dmowska, in addition to being
a regular and important scientific collaborator, is an excellent influence on Jim. Renata is
an enthusiastic polymath with a warm and cheerful personality and a seemingly endless
array of interests. She insists that he take much-deserved breaks from his research to attend
concerts, to visit art museums, to travel, to read literature, and to socialize with their large
circle of friends. JRR is increasingly active in his research, full of curiosity, creativity and
persistence. As his students can attest, he is also an excellent teacher in the classroom. He
gives lectures in a humorous, but comprehensive way that can be easily digested by his
audience. As a thesis advisor, he defines the scope of a research area in which he sees the
potential for advancement. He inspires and encourages, but does not push his students.
When a student heads in a wrong direction or reaches a dead end, he wastes no time to steer
him or her back to the right track. His good qualities as an advisor were recognized by his
recent Excellence in Mentoring Award conferred by the Graduate Student Council of
Harvard University in April 1999.
JRR recently returned from his full year sabbatical leave (January 1999 to January
2000) in Paris, France, working in the Dpartement Terre Atmosphre Ocan of cole
Normale Suprieure, and also part time at cole Polytechnique in Paliseau. His flow of
publications shows no sign of diminishing and his friends and colleagues surely will hope
that the short legend J. R. Rice will appear again and again in the scientific literature for
many years to come.

TZE-JER CHUANG JOHN W. RUDNICKI


Gaithersburg, MD Evanston, IL
List of Publications by James R. Rice

1. G. C. Sih and J. R. Rice, The Bending of Plates of Dissimilar Materials with


Cracks, Journal of Applied Mechanics, 31, (1964), pp. 477-482.
2. J. R. Rice and E. J. Brown, Discussion of Random Fatigue Failure of a Multiple
Load Path Redundant Structure by Heller, Heller and Freudenthal, in Fatigue:
An Interdisciplinary Approach (eds. J. Burke, N. Reed and V. Weiss), Syracuse
University Press, (1974), pp. 202-206.
3. J. R. Rice and F. P. Beer, On the Distribution of Rises and Falls in a Continuous
Random Process, Transactions ASME (Journal of Basic Engineering), 87D,
(1965), pp. 398-404.
4. J. R. Rice and G. C. Sih, Plane Problems of Cracks in Dissimilar Materials,
Journal of Applied Mechanics, 32, (1965), pp. 418-423.
5. J. R. Rice, F. P. Beer and P. C. Paris, On the Prediction of Some Random
Loading Characteristics Relevant to Fatigue, in Acoustical Fatigue in Aerospace
Structures (eds. W. Trapp and D. Forney), Syracuse University Press, (1965), pp.
121-144.
6. J. R. Rice, Starting Transients in the Response on Linear Systems to Stationary
Random Loadings, Journal of Applied Mechanics, 32, (1965) pp. 200-201.
7. J. R. Rice, Plastic Yielding at a Crack Tip, in Proceedings of the 1st
International Conference on Fracture, Sendai, 1965 (eds. T. Yokobori, T.
Kawasaki, and J. L. Swedlow), Vol. I, Japanese Society for Strength and Fracture
of Materials, Tokyo, (1966), pp. 283-308.
8. J. R. Rice, An Examination of the Fracture Mechanics Energy Balance from the
Point of View of Continuum Mechanics, in Proceedings of the 1st International
Conference on Fracture, Sendai, 1965 (eds. T. Yokobori, T. Kawasaki, and J. L.
Swedlow),Vol. I, Japanese Society for Strength and Fracture of Materials, Tokyo,
(1966) pp. 309-340.
9. J. R. Rice and F. P. Beer, First Occurrence Time of High Level Crossings in a
Continuous Random Process, Journal of the Acoustical Society of America, 39,
(1966) pp. 323-335.
10. J. R. Rice, Contained Plastic Deformation Near Cracks and Notches Under
Longitudinal Shear, International Journal of Fracture Mechanics, 2, (1966) pp.
426-447.
11. J. R. Rice and D. C. Drucker, Energy Changes in Stressed Bodies due to Void
and Crack Growth, International Journal of Fracture Mechanics, 3, (1967) pp.
19-27.
12. J. R. Rice, Stresses due to a Sharp Notch in a Work Hardening Elastic-Plastic
Material Loaded by Longitudinal Shear, Journal of Applied Mechanics, 34,
(1967), pp. 287-298.
xxviii LIST OF PUBLICATIONS BY J. R. RICE

13. J. R. Rice, The Mechanics of Crack Tip Deformation and Extension by Fatigue,
in Fatigue Crack Propagation, Special Technical Publication 415, ASTM,
Philadelphia, (1967), pp. 247-311.
14. J. R. Rice, Discussion of Stresses in an Infinite Strip Containing a Semi-Infinite
Crack by W.G. Knauss, Journal of Applied Mechanics, 34, (1967), pp. 248-250.
15. J. R. Rice, A Path Independent Integral and the Approximate Analysis of Strain
Concentration by Notches and Cracks, Journal of Applied Mechanics, 35, (1968),
pp. 379-386.
16. J. R. Rice and G. F. Rosengren, Plane Strain Deformation Near a Crack in a
Power Law Hardening Material, Journal of the Mechanics and Physics of Solids,
16, (1968), pp. 1-12.
17. J. R. Rice, The Elastic-Plastic Mechanics of Crack Extension, International
Journal of Fracture Mechanics, 4, (1968), pp. 41-49 (also published in
International Symposium on Fracture Mechanics, Wolters-Noordhoff Publ.,
Groningen, 1968,41-49).
18. J. R. Rice, Mathematical Analysis in the Mechanics of Fracture, Chapter 3 of
Fracture: An Advanced Treatise (Vol. 2, Mathematical Fundamentals) (ed. H.
Liebowitz), Academic Press, N.Y., (1968), pp. 191-311.
19. J. R. Rice and N. Levy, Local Heating by Plastic Deformation at a Crack Tip,
in Physics of Strength and Plasticity (ed. A. S. Argon), M.I.T. Press, Cambridge,
Mass., (1969), pp. 277-293.
20. J. R. Rice and D. M. Tracey, On the Ductile Enlargement of Voids in Triaxial
Stress Fields, Journal of the Mechanics and Physics of Solids, 17, (1969), pp.
201-217.
21. D. C. Drucker and J. R. Rice, Plastic Deformation on Brittle and Ductile
Fracture, Engineering Fracture Mechanics, 1, (1970), pp. 577-602.
22. J. Kestin and J. R. Rice, Paradoxes in the Application of Thermodynamics to
Strained Solids, in A Critical Review of Thermodynamics (eds. E.G. Stuart, B.
Gal-Or and A.J. Brainard), Mono Book Corp., Baltimore, MD (1970), pp. 275-
298.
23. H. D. Hibbitt, P. V. Marcal and J. R. Rice, A Finite Element Formulation for
Problems of Large Strain and Large Displacement, International Journal of Solids
and Structures, 6, (1970), pp. 1069-1086.
24. J. R. Rice, On the Structure of Stress-Strain Relations for Time-Dependent Plastic
Deformation in Metals, Journal of Applied Mechanics, 37, (1970), pp. 728-737.
25. J. R. Rice and M. A. Johnson, The Role of Large Crack Tip Geometry Changes
in Plane Strain Fracture, in Inelastic Behavior of Solids (eds. M. F. Kanninen, et
al.), McGraw-Hill, N.Y., (1970), pp. 641-672.
26. N. Levy, P. V. Marcal, W. J. Ostergren and J. R. Rice, Small Scale Yielding Near
a Crack in Plane Strain: A Finite Element Analysis, International Journal of
Fracture Mechanics, 7, (1971), pp. 143-156.
LIST OF PUBLICATIONS BY J. R. RICE xxix

27. J. R. Rice and N. Levy, The Part-Through Surface Crack in an Elastic Plate,
Journal of Applied Mechanics, 39, (1972), pp. 185-194.
28. N. Levy, P. V. Marcal and J. R. Rice, Progress in Three-Dimensional Elastic-
Plastic Stress Analysis for Fracture Mechanics, Nuclear Engineering and Design,
17, (1971), pp. 64-75.
29. J. R. Rice, Inelastic Constitutive Relations for Solids: An Internal Variable
Theory and Its Application to Metal Plasticity, Journal of the Mechanics and
Physics of Solids, 19, (1971), pp. 433-455.
30. J. R. Rice, Some Remarks on Elastic Crack Tip Stress Fields, International
Journal of Solids and Structures, 8, (1972), pp. 571-578.
31. J. R. Rice and D. M. Tracey, Computational Fracture Mechanics, in Numerical
and Computer Methods in Structural Mechanics (eds. S. J. Fenves et al.),
Academic Press, N.Y., (1973), pp. 585-623.
32. B. Budiansky and J. R. Rice, Conservation Laws and Energy-Release Rates,
Journal of Applied Mechanics, 40, (1973), pp. 201-203.
33. J. R. Rice and M. A. Chinnery, On the Calculation of Changes in the Earths
Inertia Tensor due to Faulting, Geophysical Journal of the Royal Astronomical
Society, 29, (1972), pp. 79-90.
34. R. J. Bucci, P. C. Paris, J. D. Landes and J. R. Rice, J Integral Estimation
Procedures, in Fracture Toughness, Special Technical Publication 514, Part 2,
ASTM, Philadelphia, (1972), pp. 40-69.
35. J. R. Rice, The Line Spring Model for Surface Flaws, in The Surface Crack:
Physical Problems and Computational Solutions (ed. J.L. Swedlow), ASME,
N.Y., (1972), pp. 171-185.
36. R. Hill and J. R. Rice, Constitutive Analysis of Elastic/Plastic Crystals at
Arbitrary Strain, Journal of the Mechanics and Physics of Solids, 20, (1972), pp.
401-413.
37. J. R. Rice, Elastic-Plastic Fracture Mechanics (Remarks for Round Table
Discusison on Fracture at the 13th International Congress of Theoretical and
Applied Mechanics, Moscow, 1972), Engineering Fracture Mechanics, 5, (1973),
pp. 1019-1022.
38. A. C. Palmer and J. R. Rice, The Growth of Slip Surfaces in the Progressive
Failure of Overconsolidated Clay, Proceedings of the Royal Society of London,
A 332, (1973), pp. 527-548.
39. J. R. Rice, Plane Strain Slip Line Theory for Anisotropic Rigic/Plastic Materials,
Journal of the Mechanics and Physics of Solids, 21, (1973), pp. 63-74.
40. J. R. Rice, P. C. Paris and J. G. Merkle, Some Further Results of J-Integral
Analysis and Estimates, in Progress in Flaw Growth and Fracture Toughness
Testing, Special Tech. Publication 536, ASTM, Philadelphia, PA (1973), pp. 231-
245.
xxx LIST OF PUBLICATIONS BY J. R. RICE

41. R. Hill and J. R. Rice, Elastic Potentials and the Structure of Inelastic
Constitutive Laws, SIAM Journal of Applied Mathematics, 25, (1973), pp. 448-
461.
42. J. R. Rice, Continuum Plasticity in Relation to Microscale Deformation
Mechanisms, in Metallurgical Effects at High Strain Rate (eds. R.W. Rohde et
al.), Plenum Press, (1973), pp. 93-106.
43. J. R. Rice, Elastic-Plastic Models for Stable Crack Growth, in Mechanics and
Mechanisms of Crack Growth (ed. M.J. May), British Steel Corporation Physical
Metallurgy Centre Publication, April 1973 (issued 1975), pp. 14-39.
44. T.- J. Chuang and J. R. Rice, The Shape of Intergranular Creep Cracks Growing
by Surface Diffusion, Acta Metallurgica, 21, (1973), pp. 1625-1628.
45. R. O. Ritchie, J. F. Knott and J. R. Rice, On the Relationship Between Critical
Tensile Stress and Fracture Toughness in Mild Steel, Journal of the Mechanics
and Physics of Solids, 21, (1973), pp. 395-410.
46. L. B. Freund and J. R. Rice, On the Determination of Elastodynamic Crack Tip
Stress Fields, International Journal of Solids and Structures, 10, (1974), pp. 411-
417.
47. J. R. Rice, Limitations to the Small Scale Yielding Approximation for Crack Tip
Plasticity, Journal of the Mechanics and Physics of Solids, 22, (1974), pp. 17-26.
48. J. R. Rice and R. M. Thomson, Ductile vs. Brittle Behavior of Crystals,
Philosophical Magazine, 29, (1974), 73-97.
49. J. R. Rice, The Initiation and Growth of Shear Bands, in Plasticity and Soil
Mechanics (edited by A. C. Palmer), Cambridge University Engineering
Department, Cambridge, (1973) pp. 263-274.
50. J. C. Nagtegaal, D. M. Parks and J. R. Rice, On Numerically Accurate Finite
Element Solutions in the Fully Plastic Range, Computer Methods in Applied
Mechanics and Engineering, 4, (1974) pp. 153-177.
51. J. R. Rice, Continuum Mechanics and Thermodynamics of Plasticity in Relation
to Microscale Deformation Mechanisms, Chapter 2 of Constitutive Equations in
Plasticity (ed. A. S. Argon), M.I.T. Press, (1975), pp. 23-79.
52. R. M. McMeeking and J. R. Rice, Finite-Element Formulations for Problems of
Large Elastic-Plastic Deformation, International Journal of Solids and Structures,
11, (1975), pp. 601-616.
53. J. R. Rice, On the Stability of Dilatant Hardening for Saturated Rock Masses,
Journal of Geophysical Research, 80, (1975), pp. 1531-1536.
54. J. R. Rice, Discussion of The Path Independence of the J-Contour Integral by G.
G. Chell and P. T. Heald, International Journal of Fracture, 11, (1975), pp. 352-
353.
55. J. W. Rudnicki and J. R. Rice, Conditions for the Localization of Deformation in
Pressure-Sensitive Dilatant Materials, Journal of the Mechanics and Physics of
Solids, 23, (1975) pp. 371-394.
LIST OF PUBLICATIONS BY J. R. RICE xxxi

56. S. Storen and J. R. Rice, Localized Necking in Thin Sheets, Journal of the
Mechanics and Physics of Solids, 23, (1975), pp. 421-441.
57. J. R. Rice, Some Mechanics Research Topics Related to the Hydrogen
Embrittlement of Metals (discussion appended to paper by J. P. Hirth and H. H.
Johnson); Corrosion, 32, (1976), pp. 22-26.
58. J. R. Rice and M. P. Cleary, Some Basic Stress-Diffusion Solutions for Fluid-
Saturated Elastic Porous Media with Compressible Constituents, Reviews of
Geophysics and Space Physics, 14, (1976), pp. 227-241.
59. J. R. Rice, Hydrogen and Interfacial Cohesion, in Effect of Hydrogen on
Behavior of Materials (eds. A.W. Thompson and I.M. Bernstein), Metallurgical
Society of AIME, (1976), pp. 455-466.
60. L.B. Freund, D.M. Parks and J. R. Rice, Running Ductile Fracture in a
Pressurized Line Pipe, in Mechanics of Crack Growth, Special Technical
Publication 590, ASTM, Philadephia, (1976), pp. 243-262.
61. J. R. Rice, The Localization of Plastic Deformation, in Theoretical and Applied
Mechanics (Proceedings of the 14th International Congress on Theoretical and
Applied Mechanics, Delft, 1976, ed. W.T. Koiter), Vol. 1, North-Holland
Publishing Co., (1976), 207-220.
62. J. R. Rice and D. A. Simons, The Stabilization of Spreading Shear Faults by
Coupled Deformation-Diffusion Effects in Fluid-Infiltrated Porous Materials,
Journal of Geophysical Research, 81, (1976), pp. 5322-5334.
63. J. R. Rice, Elastic-Plastic Fracture Mechanics, in The Mechanics of Fracture
(ed. F. Erdogan), Applied Mechanics Division (AMD) Volume 19, American
Society of Mechanical Engineers, New York, (1976), pp. 23-53.
64. J. R. Rice, Mechanics Aspects of Stress Corrosion Cracking and Hydrogen
Embrittlement, in Stress Corrosion Cracking and Hydrogen Embrittlement of
Iron Base Alloys_ (eds. R. W. Staehle et al.), National Association of Corrosion
Engineers, Houston, (1977), pp. 11-15.
65. A. P. Kfouri and J. R. Rice, Elastic/Plastic Separation Energy Rate for Crack
Advance in Finite Growth Steps, in Fracture 1977 (eds. D.M.R. Taplin et al.),
Vol. 1, Solid Mechanics Division Publication, University of Waterloo, Canada,
(1977), pp. 43-59.
66. R. J. Asaro and J. R. Rice, Strain Localization in Ductile Single Crystals, Journal
of the Mechanics and Physics of Solids, 25, (1977), pp. 309-338.
67. J. R. Rice, Pore Pressure Effects in Inelastic Constitutive Formulations for
Fissured Rock Masses, in Advances in Civil Engineering Through Engineering
Mechanics (Proceedings of 2nd ASCE Engineering Mechanics Division Specialty
Conference, Raleigh, N.C., 1977), American Society of Civil Engineers, New
York, (1977), pp. 295-297.
xxxii LIST OF PUBLICATIONS BY J. R. RICE

68. J. R. Rice, Fracture Mechanics Model for Slip Surface Propagation in Soil and
Rock Masses, in Advances in Civil Engineering Through Engineering Mechanics
(Proceedings of 2nd ASCE Engineering Mechanics Division Specialty Conference,
Raleigh, N.C., 1977), American Society of Civil Engineers, New York, NY
(1977), pp. 373-376.
69. J. R. Rice, J. W. Rudnicki and D. A. Simons, Deformation of Spherical Cavities
and Inclusions in Fluid-Infiltrated Elastic Materials, International Journal of
Solids and Structures, 14, (1978), pp. 289-303.
70. A. Needleman and J. R. Rice, Limits to Ductility Set by Plastic Flow
Localization, in Mechanics of Sheet Metal Forming (Proceedings of General
Motors Research Laboratories Symposium, October 1977, ed. D.P. Koistinen and
N.-M. Wang), Plenum Press, (1978), pp. 237-267.
71. J. R. Rice, Some Computational Problems in Elastic-Plastic Crack Mechanics,
in Numerical Methods in Fracture Mechanics (Proceedings of the First
International Conference on Numerical Methods in Fracture Mechanics, Swansea,
Wales, 1978; eds. A. R. Luxmoore and D. R. J. Owen), Department of Civil
Engineering, University College of Swansea, Wales, (1978), pp. 434-449.
72. J. R. Rice, Thermodynamics of the Quasi-Static Growth of Griffith Cracks,
Journal of the Mechanics and Physics of Solids, 26, (1978) pp. 61-78.
73. J. R. Rice and E. P. Sorensen, Continuing Crack Tip Deformation and Fracture
for Plane-Strain Crack Growth in Elastic-Plastic Solids, Journal of the Mechanics
and Physics of Solids, 26, (1978), pp. 163-186.
74. B. Budiansky and J. R. Rice, On the Estimation of a Crack Fracture Parameter by
Long-Wavelength Scattering, Journal of Applied Mechanics, 45, (1978), pp. 453-
454.
75. V. N. Nikolaevskii and J. R. Rice, Current Topics in Non-elastic Deformation of
Geological Materials, in High-Pressure Science and Technology: Sixth AIRAPT
Conference, Volume 2: Applications and Mechanical Properties (ed. K.D.
Timmerhaus and M.S. Barber), Plenum Press, New York, NY (1979), pp. 455-
464.
76. J. R. Rice, Theory of Precursory Processes in the Inception of Earthquake
Rupture, in Proceedings of the Symposium on Physics of Earthquake Sources (at
General Assembly of International Association of Seismology and Physics of the
Earths Interior, Durham, England, August 1977), Gerlands Beitrage zur
Geophysik, 88, (1979), pp. 91-127.
77. J. R. Rice and J. W. Rudnicki, Earthquake Precursory Effects due to Pore Fluid
Stabilization of a Weakening Fault Zone, Journal of Geophysical Research, 84,
(1979), pp. 2177-2193.
78. J. B. Walsh and J. R. Rice, Local Changes in Gravity Resulting from
Deformation, Journal of Geophysical Research, 84, (1979) pp. 165-170.
LIST OF PUBLICATIONS BY J. R. RICE xxxiii

79. T.- J. Chuang, K. I. Kagawa, J. R. Rice and L. B. Sills, Non-equilibrium Models


for Diffusive Cavitation of Grain Interfaces, Acta Metallurgica, Overview Paper
No. 2, 27, (1979), pp. 265-284.
80. J. R. Rice, R. M. McMeeking, D. M. Parks and E. P. Sorensen, Recent Finite
Element Studies in Plasticity and Fracture Mechanics, in Proceedings of the
FENOMECH '78 Conference (Stuttgart, edited by K.S. Pister et al.), North-
Holland Publ. Co., Vol. 2, (1979), pp. 411-442; also, Computer Methods in
Applied Mechanics and Engineering, 17/18, (1979), pp. 411-442.
81. W. Kohn and J. R. Rice, Scattering of Long Wavelength Elastic Waves form
Localized Defects in Solids, Journal of Applied Physics, 50, (1979), pp. 3346-
3353.
82. J. R. Rice, The Mechanics of Quasi-static Crack Growth, in Proceedings of the
8th U.S. National Congress of Applied Mechanics (at U.C.L.A., June 1978; ed. R.
E. Kelly), Western Periodicals Co., North Hollywood, California, (1979), pp. 191-
216.
83. B. Budiansky and J. R. Rice, An Integral Equation for Dynamic Elastic Response
of an Isolated 3-D Crack, Wave Motion, 1, (1979), pp. 187-192.
84. J. R. Rice, Plastic Creep Flow Processes in Fracture at Elevated Temperature,
in Time-Dependent Fracture of Materials at Elevated Temperature (ed. S.M.
Wolf), U.S. Department of Energy Report CONF 790236 UC-25 (June 1979), pp.
130-145.
85. B. Budiansky, D. C. Drucker, G. S. Kino and J. R. Rice, The Pressure Sensitivity
of a Clad Optical Fiber, Applied Optics, 18, (1979), pp. 4085-4088.
86. B. Cotterell and J. R. Rice, Slightly Curved or Kinked Cracks, International
Journal of Fracture, 16, (1980), pp. 155-169.
87. A. G. Evans, J. R. Rice and J. P. Hirth, The Suppression of Cavity Formation in
Ceramics: Prospects for Superplasticity, Journal of the American Ceramic
Society, 63, (1980), pp. 368-375.
88. J. R. Rice, The Mechanics of Earthquake Rupture, in Physics of the Earths
Interior (Proc. International School of Physics Enrico Fermi, Course 78, 1979;
(ed. A. M. Dziewonski and E. Boschi), Italian Physical Society and North-Holland
Publ. Co., (1980), pp. 555-649.
89. J. R. Rice, Discussion on Outstanding Problems in Geodynamics: Mechanisms
of Faulting"', in Physics of the Earths Interior (Proc. International School of
Physics Enrico Fermi, Course 78, 1979; ed. A. M. Dziewonski and E. Boschi),
Italian Physical Society and North-Holland Publ. Co., (1980) pp. 713-716.
90. J. R. Rice and J. W. Rudnicki, A Note on Some Features of the Theory of
Localization of Deformation, International Journal of Solids and Structures, 16,
(1980), pp. 597-605.
91. H. Riedel and J. R. Rice, Tensile Cracks in Creeping Solids, in Fracture
Mechanics: Twelfth Conference (ed. P.C. Paris), Special Technical Publication
700, ASTM, Philadelphia, (1980), pp. 112-130.
xxxiv LIST OF PUBLICATIONS BY J. R. RICE

92. J. R. Rice, W. J. Drugan and T. L. Sham, Elastic-Plastic Analysis of Growing


Cracks, in Fracture Mechanics: Twelfth Conference (ed. P. C. Paris), Special
Technical Publication 700, ASTM, Philadelphia, PA (1980), pp. 189-221.
93. A. Needleman and J. R. Rice, Plastic Creep Flow Effects in the Diffusive
Cavitation of Grain Boundaries, Acta Metallurgica, Overview Paper No. 9, 28,
(1980), pp. 1315-1332.
94. J. P. Hirth and J. R. Rice, On the Thermodynamics of Adsorption at Interfaces as
it Influences Decohesion, Metallurgical Transactions, 11A, (1980), pp. 1501-
1511.
95. L. Hermann and J. R. Rice, Comparison of Theory and Experiment for Elastic-
Plastic Plane-Strain Crack Growth, Metal Science, 14, (1980), pp. 285-291.
96. J. R. Rice, Pore-Fluid Processes in the Mechanics of Earthquake Rupture, in
Solid Earth Geophysics and Geotechnology (ed. S. Nemat-Nasser), American
Society of Mechanical Engineers, Appl. Mech. Div. Volume 42, New York, NY
(1980), pp. 81-89.
97. J. R. Rice, Elastic Wave Emission from Damage Processes, Journal of
Nondestructive Evaluation, 1, (1980), pp. 215-224.
98. J. R. Rice and T.- J. Chuang, Energy Variations in Diffusive Cavity Growth,
Journal of the American Ceramic Society, 64, (1981), pp. 46-53.
99. J. R. Rice, Creep Cavitation of Grain Interfaces, in Three-Dimensional
Constitutive Relations and Ductile Fracture (ed. S. Nemat-Nasser), North-Holland
Publ. Co., (1981), pp. 173-184.
100. J. R. Rice, Constraints on the Diffusive Cavitation of Isolated Grain Boundary
Facets in Creeping Polycrystals, Acta Metallurgica, 29, (1981), pp. 675-681.
101. F. K. Lehner, V. C. Li and J. R. Rice, Stress Diffusion along Rupturing Plate
Boundaries, Journal of Geophysical Research, 86, (1981), pp. 6155-6169.
102. J. R. Rice, Elastic-Plastic Crack Growth, in Mechanics of Solids: The Rodney
Hill 60th Anniversary Volume (ed. H.G. Hopkins and M.J. Sewell), Pergamon
Press, Oxford and New York, (1982), pp. 539-562.
103. W. J. Drugan, J. R. Rice and T.-L. Sham, Asymptotic Analysis of Growing Plane
Strain Tensile Cracks in Elastic-Ideally Plastic Solids, Journal of the Mechanics
and Physics of Solids, 30, 1982, pp. 447-473; erratum, 31, (1983), p. 191.
104. J. R. Rice and A. L. Ruina, Stability of Steady Frictional Slipping, Journal of
Applied Mechanics, 50, (1983), pp. 343-349.
105. J. Pan and J. R. Rice, Rate Sensitivity of Plastic Flow and Implications for Yield
Surface Vertices, International Journal of Solids and Structures, 19, (1983), pp.
973-987.
106. V. C. Li and J. R. Rice, Pre-seismic Rupture Progression and Great Earthquake
Instabilities at Plate Boundaries, Journal of Geophysical Research, 88, (1983), pp.
4231-4246.
LIST OF PUBLICATIONS BY J. R. RICE xxxv

107. V. C. Li and J. R. Rice, "Precursory Surface Deformation in Great Plate Boundary


Earthquake Sequences", Bulletin of the Seismological Society of America, 73,
(1983), pp. 1415-1434
108. J. R. Rice and J.-c. Gu, "Earthquake Aftereffects and Triggered Seismic
Phenomena", Pure and Applied Geophysics, 121, (1983), pp. 187-219.
109. J. R. Rice, "Constitutive Relations for Fault Slip and Earthquake Instabilities",
Pure and Applied Geophysics, 121, (1983), pp. 443-475.
110. J. R. Rice, "On the Theory of Perfectly Plastic Anti-Plane Straining", Mechanics
of Materials, 3, (1984), pp. 55-80.
111. W. J. Drugan and J. R. Rice, "Restrictions on Quasi-Statically Moving Surfaces
of Strong Discontinuity in Elastic-Plastic Solids", in Mechanics of Material
Behavior (the D.C. Drucker anniversary volume, ed. G.J. Dvorak and R.T. Shield),
Elsevier, (1984), pp. 59-73.
112. J. R. Rice, "Shear Instability in Relation to the Constitutive Description of Fault
Slip", in Rockbursts and Seismicity in Mines (ed. N.C. Gay and E.H. Wainwright),
Symp. Ser. No. 6, S. African Inst. Mining and Metallurgy, Johannesburg, (1984),
pp. 57-62.
113. J.-c. Gu, J. R. Rice, A. L. Ruina and S.T. Tse, "Slip Motion and Stability of a
Single Degree of Freedom Elastic System with Rate and State Dependent
Friction", Journal of the Mechanics and Physics of Solids, 32, (1984), pp. 167-
196.
114. J. R. Rice, "Comments on 'On the Stability of Shear Cracks and the Calculation of
Compressive Strength' by J.K. Dienes", Journal of Geophysical Research, 89,
(1984), pp. 2505-2507.
115. J. R. Rice, "Shear Localization, Faulting and Frictional Slip: Discussers Report",
in Mechanics of Geomaterials (Proc. IUTAM W. Prager Symp., Sept. 1983, ed.
Z.P. Bazant), J. Wiley and Sons Ltd., (1985), Chp. 11, pp. 211-216.
116. J. R. Rice, "Conserved Integrals and Energetic Forces", in Fundamentals of
Deformation and Fracture (Eshelby Memorial Symposium), ed. B.A. Bilby, K.J.
Miller and J.R. Willis, Cambridge Univ. Press, (1985) pp. 33-56.
117. P. M. Anderson and J. R. Rice, "Constrained Creep Cavitation of Grain Boundary
Facets", Acta Metallurgica, 33, (1985), pp. 409-422.
118. J. R. Rice, "First Order Variation in Elastic Fields due to Variation in Location of
a Planar Crack Front", Journal of Applied Mechanics, 52, (1985), pp. 571-579.
119. S. T. Tse, R. Dmowska and J. R. Rice, "Stressing of Locked Patches along a
Creeping Fault", Bulletin of the Seismological Society of America, 75, (1985), pp.
709-736.
120. J. R. Rice and R. Nikolic, "Anti-plane Shear Cracks in Ideally Plastic Crystals",
Journal of the Mechanics and Physics of Solids, 33, (1985), pp. 595-622.
121. J. R. Rice, "Three Dimensional Elastic Crack Tip Interactions with Transformation
Strains and Dislocations", International Journal of Solids and Structures, 21,
(1985), pp. 781-791.
xxxvi LIST OF PUBLICATIONS BY J. R. RICE

122. J. R. Rice (Editor), Solid Mechanics Research Trends and Opportunities (Report
of the Committee on Solid Mechanics Research Directions of the Applied
Mechanics Division, American Society of Mechanical Engineers), Applied
Mechanics Reviews, 38, (1985), pp. 1247-1308; published simultaneously as
AMD-Vol. 70, ASME Book No. I00198.
123. J. R. Rice, "Fracture Mechanics", in Solid Mechanics Research Trends and
Opportunities, ed. J. R. Rice, Applied Mechanics Reviews, 38, (1985), pp. 1271-
1275; published simultaneously in AMD-Vol. 70, ASME Book No. I00198.
124. J. R. Rice and S. T. Tse, "Dynamic Motion of a Single Degree of Freedom System
following a Rate and State Dependent Friction Law", Journal of Geophysical
Research, 91, (1986), pp. 521-530.
125. R. Dmowska and J. R. Rice, "Fracture Theory and Its Seismological Applications",
in Continuum Theories in Solid Earth Physics (Vol. 3 of series "Physics and
Evolution of the Earth's Interior"; ed. R. Teisseyre), Elsevier and Polish Scientific
Publishers, (1986), pp. 187-255.
126. H. Gao and J. R. Rice, "Shear Stress Intensity Factors for a Planar Crack with
Slightly Curved Front", Journal of Applied Mechanics, 53, (1986), pp. 774-778.
127. S. T. Tse and J. R. Rice, "Crustal Earthquake Instability in Relation to the Depth
Variation of Frictional Slip Properties", Journal of Geophysical Research, 91,
(1986), pp. 9452-9472.
128. P. M. Anderson and J. R. Rice, "Dislocation Emission from Cracks in Crystals or
Along Crystal Interfaces", Scripta Metallurgica, 20, (1986), pp. 1467-1472.
129. J.-S. Wang, P.M. Anderson and J. R. Rice, "Micromechanics of the Embrittlement
of Crystal Interfaces", in Mechanical Behavior of Materials - V (Proceedings of
the 5th International Conference, Beijing, 1987; ed. M.G. Yan, S.H. Zhang and
Z.M. Zheng), Pergamon Press, (1987), pp. 191-198.
130. J. R. Rice, "Mechanics of Brittle Cracking of Crystal Lattices and Interfaces", in
Chemistry and Physics of Fracture (proceedings of a 1986 NATO Advanced
Research Workshop; edited by R.M. Latanision and R.H. Jones), Martinus Nijhoff
Publishers, Dordrecht, (1987), pp. 22-43.
131. P. M. Anderson and J. R. Rice, "The Stress Field and Energy of a Three-
Dimensional Dislocation Loop at a Crack Tip", Journal of the Mechanics and
Physics of Solids, 35, (1987), pp. 743-769.
132. H. Gao and J. R. Rice, "Somewhat Circular Tensile Cracks", International Journal
of Fracture, 33, (1987), 155-174.
133. J. R. Rice, "Two General Integrals of Singular Crack Tip Deformation Fields",
Journal of Elasticity, 20, (1988), pp. 131-142.
134. H. Gao and J. R. Rice, "Nearly Circular Connections of Elastic Half Spaces",
Journal of Applied Mechanics, 54, (1987) pp. 627-634.
135. R. Hill and J. R. Rice, "Discussion of 'A Rate-Independent Constitutive Theory for
Finite Inelastic Deformation' by M.M. Carroll", Journal of Applied Mechanics, 54,
(1987), pp. 745-747.
LIST OF PUBLICATIONS BY J. R. RICE xxxvii

136. V. C. Li and J. R. Rice, "Crustal Deformation in Great California Earthquake


Cycles", Journal of Geophysical Research, 92, (1987), pp. 11,533-11,551.
137. J. R. Rice, "Tensile Crack Tip Fields in Elastic-Ideally Plastic Crystals",
Mechanics of Materials, 6, (1987), pp. 317-335.
138. J. W. Hutchinson, M. E. Mear and J. R. Rice, "Crack Paralleling an Interface
Between Dissimilar Materials", Journal of Applied Mechanics, 54, (1987), pp. 828-
832.
139. J. R. Rice and M. Saeedvafa, "Crack Tip Singular Fields in Ductile Crystals with
Taylor Power-Law Hardening, I: Anti-Plane Shear", Journal of the Mechanics and
Physics of Solids, 36, (1988), pp. 189-214.
140. J. R. Rice, "Elastic Fracture Mechanics Concepts for Interfacial Cracks", Journal
of Applied Mechanics, 55, (1988), pp. 98-103.
141. R. Dmowska, J. R. Rice, L.C. Lovison and D. Josell, "Stress Transfer and Seismic
Phenomena in Coupled Subduction Zones During the Earthquake Cycle", Journal
of Geophysical Research, 93, (1988), pp. 7869-7884.
142. J. R. Rice, "Crack Fronts Trapped by Arrays of Obstacles: Solutions Based on
Linear Perturbation Theory", in Analytical, Numerical and Experimental Aspects
of Three Dimensional Fracture Processes (eds. A. J. Rosakis, K. Ravi-Chandar
and Y. Rajapakse), ASME Applied Mechanics Division Volume 91, American
Society of Mechanical Engineers, New York, (1988), pp. 175-184.
143. J. Yu and J. R. Rice, "Dislocation Pinning Effect of Grain Boundary Segregated
Solutes at a Crack Tip", in Interfacial Structure, Properties and Design (eds. M.H.
Yoo, W.A.T. Clark and C.L. Briant), Materials Research Society Proc. Vol. 122,
(1988), pp. 361-366.
144. R. Nikolic and J. R. Rice, "Dynamic Growth of Anti-Plane Shear Cracks in Ideally
Plastic Crystals", Mechanics of Materials, 7, (1988), pp. 163-173.
145. J. R. Rice, "Weight Function Theory for Three-Dimensional Elastic Crack
Analysis", in Fracture Mechanics: Perspectives and Directions (Twentieth
Symposium), Special Technical Publication 1020, eds. R. P. Wei and R. P.
Gangloff, ASTM, Philadelphia, (1989), pp. 29-57.
146. H. Gao and J. R. Rice, "Application of 3D Weight Functions - II. The Stress Field
and Energy of a Shear Dislocation Loop at a Crack Tip", Journal of the Mechanics
and Physics of Solids, 37, (1989), pp. 155-174.
147. J. R. Rice and J.-S. Wang, "Embrittlement of Interfaces by Solute Segregation",
Materials Science and Engineering, A107, (1989), pp. 23-40.
148. M. Saeedvafa and J. R. Rice, "Crack Tip Singular Fields in Ductile Crystals with
Taylor Power-Law Hardening, II: Plane Strain", Journal of the Mechanics and
Physics of Solids, 37, (1989), pp. 673-691.
149. H. Gao and J. R. Rice, "A First Order Perturbation Analysis of Crack Trapping by
Arrays of Obstacles", Journal of Applied Mechanics, 56, (1989), pp. 828-836.
150. J. R. Rice, D. E. Hawk and R. J. Asaro, "Crack Tip Fields in Ductile Crystals",
International Journal of Fracture, 42, (1990), pp. 301-321.
xxxviii LIST OF PUBLICATIONS BY J. R. RICE

151. J. R. Rice, "Summary of Studies on Crack Tip Fields in Ductile Crystals", in


Yielding, Damage, and Failure of Anisotropic Solids (ed. J. P. Boehler),
Mechanical Engineering Publications (London), (1990), pp. 49-52.
152. P. M. Anderson, J.-S. Wang and J. R. Rice, "Thermodynamic and Mechanical
Models of Interfacial Embrittlement", in Innovations in Ultrahigh-Strength Steel
Technology (eds. G. B. Olson, M. Azrin and E. S. Wright), Sagamore Army
Materials Research Conference Proceedings, Volume 34, (1990), pp. 619-649.
153. J. R. Rice, Z. Suo and J.-S. Wang, "Mechanics and Thermodynamics of Brittle
Interfacial Failure in Bimaterial Systems", in Metal-Ceramic Interfaces (eds. M.
Rhle, A. G. Evans, M. F. Ashby and J. P. Hirth), Acta-Scripta Metallurgica
Proceedings Series, Volume 4, Pergamon Press, (1990), pp. 269-294.
154. Y. Sun, J. R. Rice and L. Truskinovsky, "Dislocation Nucleation Versus Cleavage
in Ni3Al and Ni", in High-Temperature Ordered Intermetallic Alloys (eds. L. A.
Johnson, D. T. Pope and J. O. Stiegler), Materials Research Society Proc. Vol.
213, (1991) pp. 243-248.
155. G. E. Beltz and J. R. Rice, "Dislocation Nucleation Versus Cleavage Decohesion
at Crack Tips", in Modeling the Deformation of Crystalline Solids (eds. T. C.
Lowe, A. D. Rollett, P. S. Follansbee and G. S. Daehn), The Minerals, Metals and
Materials Society (TMS), Warrendale, Penna., (1991), pp. 457-480.
156. H. Gao, J. R. Rice and J. Lee, "Penetration of a Quasistatically Slipping Crack into
a Seismogenic Zone of Heterogeneous Fracture Resistance", Journal of
Geophysical Research, 96, (1991), 21535-21548
157. J. R. Rice, "Fault Stress States, Pore Pressure Distributions, and the Weakness of
the San Andreas Fault", in Fault Mechanics and Transport Properties in Rocks
(eds. B. Evans and T.-F. Wong), Academic Press, (1992), pp. 475-503.
158. J. R. Rice, "Dislocation Nucleation from a Crack Tip: An Analysis Based on the
Peierls Concept", Journal of the Mechanics and Physics of Solids, 40, (1992), pp.
239-271.
159. G. E. Beltz and J. R. Rice, "Dislocation Nucleation at Metal/Ceramic Interfaces",
Acta Metallurgica et Materiala, 40, Supplement, (1992), pp. s321-s331.
160. J. R. Rice, G. E. Beltz and Y. Sun, "Peierls Framework for Analysis of Dislocation
Nucleation from a Crack Tip", in Topics in Fracture and Fatigue (ed. A. S.
Argon), Springer Verlag, (1992), Chapter 1, pp. 1-58.
161. M. Saeedvafa and J. R. Rice, "Crack Tip Fields in a Material with Three
Independent Slip Systems: NiAl Single Crystal", Modelling and Simulation in
Materials Science and Engineering, 1, (1992), pp. 53-71.
162. Y. Ben-Zion, J. R. Rice and R. Dmowska, "Interaction of the San Andreas Fault
Creeping Segment with Adjacent Great Rupture Zones, and Earthquake
Recurrence at Parkfield, Journal of Geophysical Research, 98, (1993), pp. 2135-
2144.
LIST OF PUBLICATIONS BY J. R. RICE xxxix

163. J. R. Rice, "Mechanics of Solids", section of the article on "Mechanics", in


Encyclopaedia Britannica (1993 printing of the 15th edition), volume 23, pp. 734-
747 and 773, (1993).
164. J. R. Rice, "Spatio-temporal Complexity of Slip on a Fault", Journal of
Geophysical Research, 98, (1993), pp. 9885-9907.
165. Y. Sun, G. E. Beltz and J. R. Rice, "Estimates from Atomic Models of Tension-
Shear Coupling in Dislocation Nucleation from a Crack Tip", Materials Science
and Engineering A, 170, (1993), pp. 67-85.
166. Y. Ben-Zion and J. R. Rice, "Earthquake Failure Sequences Along a Cellular Fault
Zone in a 3D Elastic Solid Containing Asperity and Non-Asperity Regions",
Journal of Geophysical Research, 98, (1993), pp. 14,109-14,131.
167. J. R. Rice and G. E. Beltz, "The Activation Energy for Dislocation Nucleation at
a Crack", Journal of the Mechanics and Physics of Solids, 42, (1994), pp. 333-360.
168. J. R. Rice, Y. Ben-Zion and K.-S. Kim, "Three-Dimensional Perturbation Solution
for a Dynamic Planar Crack Moving Unsteadily in a Model Elastic Solid", Journal
of the Mechanics and Physics of Solids, 42, (1994), pp. 813-843.
169. G. Perrin and J. R. Rice, "Disordering of a Dynamic Planar Crack Front in a
Model Elastic Medium of Randomly Variable Toughness", Journal of the
Mechanics and Physics of Solids, 42, (1994), pp. 1047-1064.
170. Y. Ben-Zion and J. R. Rice, "Quasi-Static Simulations of Earthquakes and Slip
Complexity along a 2D Fault in a 3D Elastic Solid", in The Mechanical
Involvement of Fluids in Faulting, Proceedings of June 1993 National Earthquake
Hazards Reduction Program Workshop LXIII, USGS Open-File Report 94-228,
Menlo Park, CA, (1994), pp. 406-435.
171. Y. Ben-Zion and J. R. Rice, "Slip Patterns and Earthquake Populations along
Different Classes of Faults in Elastic Solids", Journal of Geophysical Research,
100, (1995), pp. 12959-12983.
172. G. Perrin, J. R. Rice and G. Zheng, "Self-healing Slip Pulse on a Frictional
Surface", Journal of the Mechanics and Physics of Solids, 43, (1995), pp. 1461-
1495.
173. P. H. Geubelle and J. R. Rice, "A Spectral Method for Three-Dimensional
Elastodynamic Fracture Problems", Journal of the Mechanics and Physics of
Solids, 43, (1995), pp. 1791-1824.
174. J. R. Rice, "Text of Timoshenko Medal Speech", in Applied Mechanics Newsletter
(ed. B. Moran), American Society of Mechanical Engineers, (Spring 1995), pp. 2-
3.
175. P. Segall and J. R. Rice, "Dilatancy, Compaction, and Slip Instability of a Fluid
Infiltrated Fault", Journal of Geophysical Research, 100, (1995), pp. 22155-22171.
176. J. R. Rice and Y. Ben-Zion, "Slip Complexity in Earthquake Fault Models",
Proceedings of the National Academy of Sciences USA, 93, (1996), pp. 3811-
3818.
xl LIST OF PUBLICATIONS BY J. R. RICE

177. R. Dmowska, G. Zheng and J. R. Rice, "Seismicity and Deformation at Convergent


Margins due to Heterogeneous Coupling", Journal of Geophysical Research, 101,
(1996), pp. 3015-3029.
178. M. A. J. Taylor, G. Zheng, J. R. Rice, W. D. Stuart and R. Dmowska, "Cyclic
Stressing and Seismicity at Strongly Coupled Subduction Zones", Journal of
Geophysical Research, 101, (1996), pp. 8363-8381.
179. G. Zheng, R. Dmowska and J. R. Rice, "Modeling Earthquake Cycles in the
Shumagins Subduction Segment, Alaska, with Seismic and Geodetic Constraints",
Journal of Geophysical Research, 101, (1996), pp. 8383-8392.
180. G. E. Beltz, J. R. Rice, C. F. Shih and L. Xia, "A Self-Consistent Model for
Cleavage in the Presence of Plastic Flow", Acta Materiala, 44, (1996), pp. 3943-
3954.
181. M. F. Linker and J. R. Rice, "Models of Postseismic Deformation and Stress
Transfer Associated with the Loma Prieta Earthquake", in U. S. Geological Survey
Professional Paper 1550-D: The Loma Prieta, California, Earthquake of October
17, 1989 - Aftershocks and Postseismic Effects, (1997), pp. D253-D275.
182. A. Cochard and J. R. Rice, "A Spectral Method for Numerical Elastodynamic
Fracture Analysis without Spatial Replication of the Rupture Event", Journal of the
Mechanics and Physics of Solids, 45, (1997), pp. 1393-1418.
183. Y. Ben-Zion and J. R. Rice, "Dynamic Simulations of Slip on a Smooth Fault in
an Elastic Solid", Journal of Geophysical Research, 102, (1997), pp. 17771-17784.
184. J. W. Morrissey and J. R. Rice, "Crack Front Waves", Journal of the Mechanics
and Physics of Solids, 46, (1998), pp. 467-487.
185. M. A. J. Taylor, R. Dmowska and J. R. Rice, "Upper-plate Stressing and
Seismicity in the Subduction Earthquake Cycle", Journal of Geophysical Research,
103, (1998), pp. 24523-24542.
186. G. Zheng and J. R. Rice, "Conditions under which Velocity-Weakening Friction
allows a Self-healing versus a Cracklike Mode of Rupture", Bulletin of the
Seismological Society of America, 88, (1998), pp. 1466-1483.
187. K. Ranjith and J. R. Rice, "Stability of Quasi-static Slip in a Single Degree of
Freedom Elastic System with Rate and State Dependent Friction", Journal of the
Mechanics and Physics of Solids, 47, (1999), pp. 1207-1218.
188. J. R. Rice, "Foundations of Solid Mechanics", in Mechanics and Materials:
Fundamentals and Linkages (eds. M. A. Meyers, R. W. Armstrong, and H.
Kirchner), Chapter 3, Wiley, in press, (1999).
189. J. W. Morrissey and J. R. Rice, "Perturbative Simulations of Crack Front Waves",
Journal of the Mechanics and Physics of Solids, in press.
List of Contributors

Professor Peter M. Anderson, Department of Materials Science and Engineering, The Ohio
State University, 2041 College Road, Columbus, OH 43210-1179 U. S. A.

Professor A. G. Atkins, Department of Engineering, University of Reading, Reading,


BG6 6AY, UK

Professor Leslie Bank-Sills, The Dreszer Fracture Mechanics Laboratory, Department of


Solid Mechanics, Materials and Structures, The Fleischman Faculty of Engineering, Tel
Aviv University, 69978 Ramat Aviv, Israel

Dr. B Blug, Fraunhofer-Institut fr Werkstoffmechanik,Whlerstr. 11,79108 Freiburg,


Germany

Dr. Vinodkumar Boniface, The Dreszer Fracture Mechanics Laboratory, Department of


Solid Mechanics, Materials and Structures, The Fleischman Faculty of Engineering, Tel
Aviv University, 69978 Ramat Aviv, Israel

Professor Allan F. Bower, Division of Engineering, Brown University, Providence, RI


02912, U.S.A.

Dr. B. Chen,. Department of Mechanical and Industrial Engineering, University of Illinois,


Urbana, IL 61801

Dr. Z. Chen, Institute of Materials Research and Engineering, 3 Research Link, Singapore
117602

Dr. W. Y. Chien, Department of Mechanical Engineering and Applied Mechanics, The


University of Michigan, Ann Arbor, MI 48109, USA

Dr.J. W. Cho, Technical Center, Deawoo Heavy Industries Co, Inchun, Korea

Dr. T.-J. Chuang, Ceramics Division, National Institute of Standards and Technology,
Gaithersburg, MD 20899-8521, U. S. A.

Dr. Brian Cotterell, Institute of Materials Research and Engineering, 3 Research Link,
Singapore 117602
x1ii LIST OF CONTRIBUTORS

Professor Walter W.Drugan, Department of Engineering Physics, University of Wisconsin,


Madison, 1500 Engineering Drive, Madison, WI 53706

Professor Glenn E. Beltz, Department of Mechanical and Environmental Engineering,


University of California, Santa Barbara, CA 93106-5070, USA

Professor Huajian Gao, Division of Mechanics and Computation, Department of


Mechanical Engineering, Stanford University, Stanford, CA 94305-4040

Mr. Anja Haug, Materials Department, University of California, Santa Barbara, California
93106 USA

Professor Young Huang, Department of Mechanical and Industrial Engineering, University


of Illinois, Urbana, IL 61801

Mr. H.-M. Huang, Department of Mechanical Engineering and Applied Mechanics, The
University of Michigan, Ann Arbor, MI 48109, USA

Professor Mark Kachanov, Department of Mechanical Engineering, Tufts University,


Medford, MA 02155

Dr. E. Karapetian, Department of Mechanical Engineering, Tufts University, Medford, MA


02155, U.S.A.

Dr.Patrick A Klein, Sandia National Laboratories, Mail Stop 9161,P.O. Box 0969,
Livermore, CA 94551

Professor Shiro Kubo, Department of Mechanical Engineering and Systems, Graduate


School of Engineering, Osaka University, 2-1, Yamadaoka, Suita, Osaka 565-0871
Japan

Dr. L. L. Fischer, Department of Mechanical and Environmental Engineering, University


of California,Santa Barbara, CA 93106-5070, USA

Dr. L. E. Levine, Maaterials Science and Engineering Lab., National Institute of Standards
and Technology, Gaithersburg, MD 20899

Professor Victor C. Li, Department of Civil and Environmental Engineering, University


of Michigan, Ann Arbor, MI, 48109-2125

Mr. W. Lu, Mechanical and Aerospace Engineering Department and Materials Institute,
Princeton University, Princeton, NJ 08544
LIST OF CONTRIBUTORS xliii

Dr. S. R. MacEwen, Alcan International Ltd., P.O. Box 8400, Kinston, Ontario, K7L 5L9,
Canada

Professor Robert M. McMeeking, Department of Mechanical and Environmental


Engineering, University of Californi, Santa Barbara, California 93106, USA

Professor Sinisa Dj. Mesarovic, Department of Materials Science and Engineering,


University of Virginia, Charlottesville, VA 22903 U. S. A.

Professor Joe Pan, Department of Mechanical Engineering and Applied Mechanics, The
University of Michigan, Ann Arbor, MI 48109, USA

Dr. Hermann Riedel, Fraunhofer-Institut fr Werkstoffmechanik,Whlerstr. 11,79108


Freiburg, Germany

Professor Asher A. Rubinstein, Department of Mechanical Engineering, Tulane University,


New Orleans, LA 70118, U. S. A.

Professor J. W .Rudnicki, Department of Civil Engineering, Northwestern University,


2145 Sheridan Road, Evanston, IL 60208-3109

Dr. I. Sevostianov, Department of Mechanical Engineering, Tufts University, Medford,


MA 02155

Dr. Y. Shim, Center for Simulational Physics, University of Georgia, Athens, GA 30602

Professor Z. Suo, Mechanical and Aerospace Engineering Department, and Materials


Institute, Princeton University, Princeton, NJ 08544

Dr. S. C. Tang, Ford Research Lab., P.O. Box 2053, MD3135/SRL, Dearborn, MI
48121, U.S.A.

Mr. Zhibo Tang, Division of Engineering, Brown University, Providence, RI 02912

Dr. Robb M. Thomson, Maaterials Science and Engineering Lab., National Institute of
Standards and Technology, Gaithersburg, MD 20899

Dr. Jian-Sheng Wang, Northwestern University, Evanston, IL 60201, USA

Dr. P. D. Wu, Alcan International Ltd., P.O. Box 8400, Kinston, Ontario, K7L 5L9,
Canada

Dr. Z. C. Xia, Ford Research Lab., P.O. Box 2053, MD3135/SRL, Dearborn, MI 48121
x1iv LIST OF CONTRIBUTORS

Dr. Xiao J. Xin, Department of Mechanical and Nuclear Engineering, Kansas State
University, 338 Rathbone Hall, Manhattan, KS 66506-5205 U. S. A.

Professor Jin Yu, Department of Materials Science and Engineering,Korea Advanced


Institute of Science and Technology, P.O. Box 201, Chongryang. Seoul, Korea
APPROXIMATE YIELD CRITERION FOR ANISOTROPIC
POROUS SHEET METALS AND ITS APPLICATIONS TO
FAILURE PREDICTION OF SHEET METALS UNDER FORMING
PROCESSES
W. Y. CHIEN, H.-M. HUANG AND J. PAN
Department of Mechanical Engineering and Applied Mechanics
The University of Michigan, Ann Arbor, MI 48109, USA

AND

S. C. TANG
Ford Research Laboratory
Ford Motor Company, Dearborn, MI 48121, USA

Abstract: An approximate anisotropic yield criterion for anisotropic sheet metals


containing spherical voids is validated using a three-dimensional finite element analysis.
An aggregate of periodically arranged cubes containing spherical voids is modeled using
a unit cell method. Hills quadratic anisotropic yield criterion is used to describe the
normal anisotropy and planar isotropy of the matrix. The metal matrix is first assumed to
be elastic perfectly plastic and incompressible. The results of the finite element analysis
can be in good agreement with those based on the proposed yield criterion by introducing
three fitting parameters in the yield criterion. This modified yield criterion is adopted in
a failure prediction methodology that can be used to determine the failure of sheet metals
under forming operations. The material imperfection approach is employed to predict
failure/plastic localization by assuming a slightly higher void volume fraction inside
randomly oriented imperfection bands. Finally, the failure prediction methodology is
applied to predict the failure of a mild steel sheet metal in a fender forming process.

1. Introduction

Structural metals usually contain some form of second-phase particles such as maganese
sulfides and carbides in steels. These particles usually provide strain concentration sites
for void nucleation, growth and coalescence that lead to ductile fracture. In order to
model the plastic flow and fracture of these ductile structural metals, Gurson (1977)
conducted an upper bound analysis of simplified models containing voids and proposed
an approximate yield criterion for porous materials where the matrices obey the von
Mises yield criterion.
1
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 115.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
2 W.Y.CHIEN, ET AL.

Based on Gursons yield criterion (1977), Yamamoto (1978) investigated plastic flow
localization with the assumption of a slightly higher initial void concentration as the
initial material imperfection inside thin planar bands in a material element of interest.
Needleman and Triantafyllidis (1978) further examined the effects of void growth based
on Gursons yield criterion (1977) on localized necking in biaxially stretched sheets and
compared the results with those from various types of initial imperfections. It was shown
in Yamamoto (1978), Needleman and Triantafyllidis (1978), Saje et al. (1982) and
Tvergaard (1981) that the porous material model based on Gursons yield criterion
predicts failure qualitatively in accord with experimental results. In order to predict sheet
metal failure quantitatively in accord with the experimental results under biaxial
stretching conditions, Mear and Hutchinson (1985) introduced a family of constitutive
laws to address the sensitivity of failure prediction to the yield surface curvature. They
considered that the evolution of yield surface follows a simple rule of a combination of
isotropic expansion and kinematic translation.
Tvergaard (1981, 1982) introduced three additional fitting parameters in Gursons yield
criterion by comparing the results of shear band instability in square arrays of cylindrical
holes and axisymmetric spherical holes based on finite element models with those based
on Gursons yield criterion. Saje et al. (1982) studied the void nucleation effects on
shear localization in rate insensitive porous plastic solids using the modified yield
criterion. A parallel work was carried out by Pan et al. (1983) with consideration of
material rate sensitivity. The modified yield criterion was also used in the analysis of the
cup-cone fracture in a round tensile bar by Tvergaard and Needleman (1984).
The matrix material in the original Gurson model was assumed to be isotropic.
However, sheet metals for stamping applications usually display certain extent of plastic
anisotropy due to cold or hot rolling processes. In general, an average value of the
anisotropy parameter R, defined as the ratio of the transverse plastic strain rate to the
through-thickness plastic strain rate under in-plane uniaxial loading conditions, is used to
characterize the sheet anisotropic plastic behavior. Graf and Hosford (1990) investigated
the effects of R on forming limit using different yield criteria. They found that when
Hills quadratic anisotropic yield criterion (1948) is employed, R has significant effects
on forming limit. When the higher order yield criterion of Hosford (1979) is employed,
R has virtually no effects on forming limit. This indicates that further research on the
effects of plastic anisotropy or the effects of constitutive laws/plastic hardening on
forming limit is needed.
In view of possible significant effects of plastic anisotropy on the forming limit in sheet
forming processes, a Gurson type of approximate yield criterion, which can be used to
account for the matrix plastic anisotropy, is needed to investigate forming limit and to
characterize ductile fracture processes in sheet metal forming applications (Liao et al.,
1997). For simplicity, sheet metals were assumed to have normal anisotropy and planar
isotropy. In order to develop an approximate macroscopic yield criterion for voided
sheet metals with normal anisotropy and planar isotropy, a simplified sheet model
containing a through-thickness hole under plane stress conditions was considered in Liao
et al. (1997). In their investigation, the matrix material was characterized by Hills
quadratic anisotropic yield criterion (1948) and Hills non-quadratic anisotropic yield
criterion (1979). Upper bound analyses were carried out and the numerical results based
on both Hills quadratic and non-quadratic anisotropic yield criteria were fitted well by a
YIELD AND FAILURE OF SHEET METALS 3

closed-form macroscopic yield criterion. An anisotropic Gurson yield criterion for sheet
metals with spherical voids was also proposed.
In order to validate the accuracy of the proposed anisotropic Gurson yield criterion by
Liao et al. (1997), a three-dimensional finite element analysis is employed for a cube
model containing a spherical void (Chien et al., 2000). The analysis is performed for
various void volume fractions as well as different R values for elastic perfectly plastic
materials. As in Tvergaard (1981, 1982), the anisotropic Gurson yield criterion is
modified by adding three fitting parameters to fit the results based on the modified yield
criterion with the finite element computational results. Finite element computations with
consideration of material strain hardening under multiaxial proportional straining
conditions are also performed. The results of finite element simulations are compared
with those based on the modified anisotropic Gurson yield criterion. Finally, the
modified anisotropic Gurson yield criterion is adopted here in a failure prediction
methodology that can be used to predict failure in anisotropic, rate-sensitive sheet metals
under forming processes. The material imperfection approach is used to predict sheet
metal failure by assuming a slightly higher void volume fraction inside randomly
oriented imperfection bands in the critical sheet element of interest. The failure of sheet
metals under forming processes is defined where the failure or plastic flow localization in
an imperfection band of the critical element is first met. Finally, conclusions are given.

2. Anisotropic Gurson Yield Criterion

Liao, Pan and Tang (1997) considered a material element of sheet metals with arbitrary
shaped voids as shown in Figure 1(a). The void volume fraction was assumed to be
small. The sheet element was assumed to be subject to in-plane loading conditions for
sheet metal forming applications. The matrix surrounding the voids was assumed to be
rigid perfectly-plastic, incompressible and rate-insensitive to take advantage of the upper
bound analysis. Liao et al. (1997) considered a simplified sheet model as shown in
Figure 1(b) where a sheet contains a periodic array of circular through-thickness holes.
A sheet cell model as shown in Figure 1(c) was then considered for upper bound
analyses. Under axisymmetric loading, a closed-form upper-bound macroscopic yield
criterion CR can be derived as

(1)

w h e r e e represents the macroscopic effective stress based on Hills quadratic


anisotropic yield criterion, o is the matrix yield stress under in-plane uniaxial loading
conditions, is the void volume fraction, R is the anisotropic parameter, and m
( = kk /3 ) is the macroscopic mean stress.
Figure 2 shows the upper bound solutions as open symbols for R = 1.8, which
represents the typical value of R for low carbon steels. The results based on the closed-
form macroscopic yield criterion in Equation (1) are also plotted as various types of
4 W.Y.CHIEN, ET AL.

curves. Note that under macroscopic pure shear loading ( m = 0), the solutions are
located along the e /o axis. For a given , the solution with the largest macroscopic
mean stress as shown in the figure represents the upper bound solution for equal biaxial
tension. The solutions between the pure shear and the equal biaxial tension are for the
other possible plane stress loading conditions. A reasonable match of the upper bound
solutions with the macroscopic yield criterion for various plane stress deformation modes
can be seen in Figure 2.

Figure 1(a). A sheet element with arbitrarily shaped voids.

Figure 1(b). A simplified sheet model with a periodic array of circular through-thickness holes.

Figure 1(c). A sheet cell for the simplified sheet model.

In order to understand the accuracy of the upper bound solutions and the closed-form
solutions, the macroscopic in-plane mean stress is also obtained for the sheet cell model
subject to a uniform radial traction under fully yielded conditions. The exact limit
YIELD AND FAILURE OF SHEET METALS 5

solutions for various values of are shown as solid symbols in Figure 2. As shown in
Figure 2, the upper bound macroscopic effective stress and the mean stress are slightly
larger than those of the exact limit solution for equal biaxial tension. The good
agreement of the exact limit solution and the approximate yield criterion under equal
biaxial tension indicates the accuracy of the closed-form upper-bound solutions and the
validity of the approximate yield criterion near equal biaxial tension.

Figure 2. Comparison of the upper bound solutions based on the sheet cell model and the yield contours of the
closed-form anisotropic Gurson yield criterion for various void volume fractions for R=1.8. The exact
solutions of the macroscopic in-plane mean stress under fully yielded conditions are also shown as bullets.

In the original Gursons work (1977) for isotropic materials, the basic of the yield
criterion for the cylindrical void model is

(2)

and the basic form of the yield criterion for the spherical void model is

(3)

It can be seen that both the yield criteria have the same form except a factor of in the
cosh function. Therefore, the closed-form solution obtained from the sheet cell model
6 W.Y.CHIEN, ET AL.

with inclusion of a factor of 1/ can be used as a first-order approximate solution for


the spherical void model with the matrix having mild normal anisotropy. An
approximate anisotropic Gurson yield criterion for the spherical void model can be
written as

(4)

3. Modified Anisotropic Gurson Yield Criterion

Since Equation (4) is again approximate in nature, it is necessary to validate the accuracy
of this yield criterion. Therefore, a three-dimensional finite element model is considered.
As in Tvergaard (1981, 1982), a modified yield criterion is proposed in Liao et al. (1997)
as

(5)

where q 1, q2 , and q 3 are parameters to fit the finite element computational results.
The finite element model considered here is a cube containing a spherical void, as
shown in Figure 3. Due to the symmetry of the cube and the applied loads, only one-
eighth of the cube is analyzed here, as shown in Figure 4.

Figure 3. A voided cube model.

Figure 4. One-eighth of the voided cube model.


YIELD AND FAILURE OF SHEET METALS 7

The boundary conditions of the model are set as follows. The bottom, front, and left
surfaces are constrained to have zero normal displacements to satisfy the symmetry
conditions. The void surface is specified to have zero traction. Uniform normal
displacements are applied on the top, back, and right faces. Pure shear, uniaxial tension
and equal biaxial tension are considered for plane stress conditions. In order to use the
one-eighth voided cube model to investigate the plastic behavior under pure in-plane
shear, a uniform normal displacement is applied to the right face in the x1 direction and
the same amount of normal displacement is applied uniformly to the back face in the
negative x2 direction to result in pure in-plane shear. A triaxial loading with high mean
stress and a pure hydrostatic tension loading are also applied to complete the analysis.
The matrix material is assumed to be elastic perfectly plastic. Poissons ratio v of the
matrix material is assumed to be 0.33 and the ratio of the matrix yield stress o to
Youngs modulus E is set at 2 10-7 . With the small value of o /E, the matrix material
can be considered as nearly rigid-perfectly plastic. Several initial void volume fractions
( = 0.01, 0.04, 0.09 and 0.12) and R = 1.6 are used here to validate the proposed yield
criterion in Equation (5).
The commercial finite element program ABAQUS (Hibbitt et al., 1998) is used to
perform the computations. Twenty node elements with a reduced integration scheme are
used here such that the model is free from overconstraint. The macroscopic yield point is
defined as the limited stress state where massive plastic deformation occurs. The
corresponding macroscopic effective stress and macroscopic mean stress are then
calculated and compared with those based on the anisotropic Gurson yield criterion in
Equation (5).

Figure 5. Comparison of the modified anisotropic Gurson yield criterion (curves) and the finite element results
(open symbols) for R = 1.6.
8 W.Y.CHIEN, ET AL.

In addition to the elastic perfectly plastic material model employed to calculate the fully
plastic limits, the macroscopic plastic flow characteristics due to the material matrix
hardening are also investigated for several proportional loading conditions (Chien et al.,
2000). The in-plane shear stress and plastic shear strain of the matrix is assumed to
follow a power-law relation. Then the matrix effective in-plane tensile stress M as a
function of the effective in-plane tensile strain M
P can be written as

(6)

where G represents the shear modulus under in-plane shear loading and N represents the
hardening exponent. Here N = 0.1 is used.

Figure 6. The stress-strain behavior of the finite element results and the continuum models with = 0.09 for
R = 1.6.

The finite element computational results are used to evaluate the accuracy of the
anisotropic Gurson yield criterion. Figure 5 shows the finite element computational
results (represented by open symbols) compared with those of the modified anisotropic
Gurson yield criterion (represented by various types of curves). The values of the fitting
parameters are selected as q1 = 1.45, q 2 = 0.9 and q 3 = 1.6 for R = 1.6. With the fitting
parameters, it can be seen that the finite element computational results are in good
agreement with those of the modified anisotropic Gurson yield criterion. It should be
noted that when the macroscopic mean stress equals zero, the macroscopic effective
stresses from our finite element computations are slightly lower than those predicted by
the modified yield criterion. The reason for the earlier yielding from our finite element
computations can be attributed to the shear localization in the matrix material. In order
to investigate further the accuracy of the modified anisotropic Gurson yield criterion, the
YIELD AND FAILURE OF SHEET METALS 9

macroscopic hardening behavior from finite element computations is compared with that
based on the modified anisotropic Gurson yield criterion. Figure 6 shows the normalized
macroscopic tensile stress 11 / o as a function of the macroscopic tensile strain E11 for
= 0.09 and R = 1.6 under uniaxial tensile conditions. As shown in the figure, the finite
element computational results agree with those of the modified anisotropic Gurson yield
criterion at low strains, and approach to those of the unmodified anisotropic Gurson
yield criterion at large strains. Similar trends are also seen in Hom and McMeeking
(1989) for isotropic materials. More computational results for various values of R under
different multiaxial proportional loading conditions can be found in Chien et al. (2000).

4. Evolution of Void Volume Fraction

Tvergaard and Needleman (1984) introduced the void volume fraction parameter ,
which is a function of the void volume fraction , into the Gurson model in order to
account for the gradual loss of stress carrying capacity due to void coalescence. The
function ( ) in Tvergaard and Needleman (1984) is

(7)

In Equation (7), the quantity u is defined as the limiting value of ( ) as the stress
carrying capacity goes to zero. c represents the critical void volume fraction, and
represents the void volume fraction at final failure. Based on the experimental studies of
Brown and Embury (1973) and Goods and Brown (1979) and the numerical analysis of
Andersson (1977), the values of c and were chosen as 0.15 and 0.25, respectively
(Tvergaard, 1982). Here, we adopt the function ( ) into the modified anisotropic
Gurson yield criterion as

(8)

In addition to the effect of void coalescence that has been addressed by the function
( ) , the evolution of void volume fraction can be related to the nucleation of voids
and growth/collapse of the existing voids. The increase/decrease of void volume fraction
due to growth/collapse can be obtained with the assumption of plastic incompressibility
of the matrix material based on the original work of Gurson (1977). For the evolution of
void volume fraction due to nucleation, two models can be considered: one is the plastic
strain controlled nucleation model suggested by Gurson (1977) based on the
experimental data of Gurland (1972). The other is the stress controlled nucleation model
in which void nucleation depends on the maximum stress transmitted across the particle-
10 W.Y.CHIEN, ET AL.

matrix interface as discussed in Argon and Im (1975). Thus, the rate of void volume
fraction can be expressed as [Chu and Needleman (1980) and Saje et al. (1982)]:

(9)

Here tr represents the macroscopic dilatational plastic strain rate, represents the
matrix average plastic strain rate, represents the matrix average stress rate, and
represents the Jaumann rate of the macroscopic mean stress. Here A and B are void
nucleation parameters for the strain-controlled and stress-controlled nucleation models,
respectively.

5. Yield Surface Curvature Effects

Based on Gursons yield criterion for isotropic porous materials, Mear and Hutchinson
(1985) considered a family of yield surfaces with different yield surface curvatures in
order to fit well with the experimental results on necking instability. Mear and
Hutchinson (1985) specified the size of the yield surface, F , which is a linear
combination of the matrix initial yield stress y and the matrix flow stress M . Here,
this yield (or potential) surface is regarded as the curvature surface. The details of
including the curvature surface into our plasticity model can be found in Huang et al.
(2000a). The tangent modulus procedure of Peirce et al. (1984) is employed to obtain the
evolution of the rate-sensitive constitutive relations. The derivation of the constitutive
relations including anisotropic hardening with consideration of void growth/collapse is
detailed in Huang et al. (2000b).

6. Failure Localization Analysis

We employ a Lagrangian formulation and take the initial undeformed configuration as


the reference. The coordinates of a material point relative to a fixed Cartesian frame in
the undeformed configuration, xi , are taken as the convected coordinates. In the current
deformed state, the coordinates of a material point, referred to the reference Cartesian
base vectors, are denoted by Latin indices range from 1 to 3 and summation
convention is adopted for repeated indices.
In the present analysis, many infinitely thin imperfection bands with different
orientations are assumed to exist in a material element of interest. Figure 7 shows a
material element having an imperfection band under plane stress conditions. In the
figure, N represents the normal vector to the band in the undeformed state. The band
angle , which is used to represent the orientation of the imperfection band, is the angle
between the direction of N and a loading direction of interest, that is taken as the x2
direction as shown in Figure 7.
YIELD AND FAILURE OF SHEET METALS 11

Figure 7. A material element having an imperfection band under plane stress conditions

Homogeneous deformations inside and outside the band are assumed to occur
throughout a deformation history. Here, a superscript or a subscript b represents a
quantity inside the band, while a superscript or a subscript o represents a quantity
outside the band.
As the deformation proceeds, compatibility requires [Hill (1962) and Rice (1976)]:

(10)

where C is a vector denoting the discontinuity across the band. Also, equilibrium
requires that the normal tractions are continuous over the band interface. Therefore, the
equilibrium equation can be given in terms of the first Piola-Kirchoff stress S and the
normal vector N as

with (11)

where T ik represent the contravariant components of the Kirchoff stress tensor T.


Combining the rate form of the compatibility equations in Equation (10) and the rate
form of the equilibrium equations in Equation (11) with the constitutive relations gives a
set of equations for
Given a prescribed deformation history outside the band, , and the initial conditions,
the set of equations for can be solved incrementally to determine the deformation
history inside the band. The condition of failure/plastic localization is reached when the
ratio of the normal component of the deformation gradient rate in the direction of N
inside the band to that outside the band becomes infinity.
A sheet metal under a fender forming process, which is the benchmark problem of the
1993 NUMISHEET conference as shown in Figure 8, has been considered in Huang et
12 W.Y.CHIEN, ET AL.

al. (2000a). The deformation history including the relative rotation of principal stretch
directions for the critical element as shown in Figure 8 has been identified. The failure of

Figure 8. The critical element and the initial major principal stretch directions of the element in a sheet metal
under a fender forming operation.

Figure 9. The principal strain history of the critical element based on the FEM fender forming simulation and
the corresponding predicted failure strains with rotating and fixed principal stretch directions. The calculated
forming limit diagram (FLD) under proportional loading conditions for the mild steel is also presented as a
solid curve for comparison.
YIELD AND FAILURE OF SHEET METALS 13

the critical element of the mild steel sheet is determined based on the growth of
microvoids in imperfection bands under the non-proportional deformation history. The
values of the initial imperfection in terms of the void volume fraction and the curvature
parameter b are selected to fit the forming limit diagram (FLD) for the mild steel under
proportional straining conditions. The values of and b are 0.000025 and 0.25,
respectively. The value of b remains the same as that in Huang et al. (2000a). The value
of f is smaller than that in Huang et al. (2000a) due to the selection of q 1 = 1.45, q 2 = 0.9
and q3 =1.6 in this investigation whereas q1 , q 2 and q 3 were set at unity in Huang et al.
(2000a). Figure 9 shows the principal strain history for the critical element in the sheet
metal based on the FEM fender forming simulation and the corresponding predicted
failure strains with rotating and fixed principal stretch directions. The calculated FLD
under proportional loading conditions for the mild steel is also presented as a solid curve
in the same figure for comparison. Here, the predicted major principal failure strain with
consideration of rotating principal stretch directions is 13.6 percent lower than that
without consideration of rotating principal stretch directions under this specific
deformation history. Note that in Huang et al. (2000a), where q1, q2 and q 3 were set at
unity, the predicted major principal failure strain with consideration of rotating principal
stretch directions is 8.2 percent lower than that without consideration of rotating
principal stretch directions under this specific deformation history. This indicates that
the selection of q1 , q 2 and q 3 different from unity does affect the prediction of failure.
The effects of rotating principal stretch directions could be more difficult to predict when
the non-proportional straining path encompasses the negative minor principal strain
range. The detailed FEM forming simulation and failure prediction results are presented
in Huang et al. (2000a).

7. Conclusions

A simplified sheet cell model was first adopted in Liao et al. (1997) to obtain upper
bound solutions for anisotropic voided sheets under plane stress conditions. In their
work, the matrix of the voided sheet was assumed to be rigid perfectly plastic,
incompressible and rate-insensitive. Hills quadratic and non-quadratic anisotropic yield
criteria were used to describe the matrix normal anisotropy. A closed-form macroscopic
yield criterion for the sheet model with a through-thickness hole was derived based on
Hills quadratic anisotropic yield criterion under axisymmetric loading conditions. The
upper bound solutions were fitted well by the closed-form macroscopic yield criterion
under plane stress loading conditions.
Based on the original Gurson models for isotropic materials, the inclusion of a factor of
in the cosh function of the yield criterion was suggested to obtain a first-order
approximation for the sheet material containing spherical voids (Liao et al., 1997). Here,
a three-dimensional finite element analysis is employed to validate the anisotropic
Gurson yield criterion. The results of the finite element calculations indicate that three
fitting parameters are needed to improve the applicability of the anisotropic Gurson yield
criterion. With appropriate selection of fitting parameters, it is found that the finite
element results can be fitted by those based on the modified anisotropic Gurson yield
14 W.Y.CHIEN, ET AL.

criterion for the range of strains and loading conditions investigated in Chien et al.
(2000).
The modified anisotropic Gurson yield criterion is adopted to predict the failure of a
sheet metal under a fender forming operation. Note that a general deformation history
for a critical element including the relative rotation of principal stretch directions was
identified in a benchmark fender forming problem (Huang et al., 2000a). The failure of
the critical sheet element is predicted based on the growth of voids in various oriented
imperfection bands under this non-proportional deformation history. The current band
orientation at failure implies the possible splitting failure direction. Here, the predicted
major principal failure strain with consideration of rotating principal stretch directions is
13.6 percent lower than that without consideration of rotating principal stretch directions
under this specific deformation history. As indicated in Huang et al. (2000b), elastic
unloading/reloading can also have significant effects on failure in forming processes.
Therefore, a more general constitutive relation that can be used to describe the material
behavior under loading/unloading conditions with consideration of damage evolution was
formulated (Huang et al., 2000b). This type of constitutive relation with consideration of
void nucleation, growth/collapse and final coalescence should be further validated and
modified by experiments. Then the constitutive relation can be implemented into
computer codes to predict the failure of sheet metals in conjunction with the FEM
simulations of forming processes.

Acknowledgement

The support of this work by Ford Motor Company is greatly appreciated.

References

Argon, A. S. and Im, J. (1975), Separation of second phase particles in spheridized 1045 steel, Cu-0.6pct Cr
alloy, and maraging steel in plastic straining, Metall. Trans., 6A, 839.
Andersson, H. (1977), Analysis of a model for void growth and coalescence ahead of a moving crack tip, J.
Mech. Phys. Solids, 25, 217.
Brown, L. M. and Emgury, J. D. (1973), Proc. 3rd Int.Conf. on Strength of Metals and Alloys, pp. 164-169, Inst.
Metalls, London.
Chien, W. Y., Pan, J and Tang, S. C. (2000), to be submitted for publication.
Chu. C.-C., Needleman, A. (1980), Void nucleation effects in biaxially stretched sheets, J. Eng. Mater. Tech.,
102, 249.
Goods, S. H. and Brown, L. M. (1979), Nucleation of cavities by plastic deformation, Acta. Metall., 27, 1.
Graf, A., Hosford, W. F. (1990), Calculations of forming limit diagrams, Metall. Trans., 21A, 87.
Gurland, J. (1972), Observations on fracture of cementite particles in a spheroidized 1.05% C steel deformed at
room temperature, Acta. Metall., 20, 735.
Gurson, A. L. (1977), Continuum theory of ductile rupture by void growth: part I yield criteria and flow rules
for porous ductile media, J. Eng. Mater. Tech., 99, 2.
Hibbitt, H. D., Karlsson, B. I. and Sorensen, E. P. (1998), ABAQUS user manual, Version 5-8.
Hill, R. (1948), A theory of the yielding and plastic flow of anisotropic metals, Roy. Soc. London Proc., 193A,
281.
Hill, R. (1962), Acceleration waves in solids, J. Mech. Phys. Solids, 10, 1.
Hill, R. (1979), Theoretical plasticity of textured aggregates, Math. Proc. Camb. Philos. Soc., 85, 179.
Hom, C. L. and McMeeking R. M. (1989), Void growth in elastic-plastic materials, J. Appl. Mech., 56, 309.
YIELD AND FAILURE OF SHEET METALS 15

Hosford, W. F. (1979), On yield loci of anisotropic cubic metals, Proc. 7 th North Am. Metalworking Res. Conf.,
SME, Dearborn, MI, p. 191.
Huang, H.-M., Pan, J. and Tang, S. C. (2000a), Failure prediction in anisotropic sheet metals under forming
operations with consideration of rotating principal stretch directions, Int. J. Plast., 16, 611.
Huang, H.-M., Pan, J. and Tang, S. C. (2000b), Failure prediction in anisotropic sheet metals containing voids
under biaxial straining conditions with prebending/unbending, to appear in Int. J. Plast..
Liao, K.-C., Pan, J. and Tang, S. C. (1997), Approximate yield criteria for anisotropic porous ductile sheet
metals, Mech. Mater., 26, 213.
Mear, M. E. and Hutchinson, J. W. (1985), Influence of yield surface curvature on flow localization in dilatant
plasticity, Mech. Mater., 4, 395.
Needleman, A. and Triantafyllidis, N. (1978), Void growth and local necking in biaxial stretched sheets, J. Eng.
Mater. Tech., 100, 164.
Marciniak, Z. and Kuczynski, K. (1967), Limit strains in the processes of stretch forming sheet metal, Int. J.
Mech. Sci., 9, 609.
Pan, J., Saje, M. and Needleman, A. (1983), Localization of deformation in rate sensitive porous plastic solids,
Int. J. Fract., 21, 261.
Peirce, D., Shih, C. F. and Needleman, A. (1984), A tangent modulus method for rate dependent solids, Comp.
& Struct., 18, 875.
Rice, J. R. (1976), Proc. 14t h Int. Cong. on Theoretical and Applied Mechanics, Ed. Koiter , W. T., 1, pp. 207-
220, Delft, North-Holland.
Saje, M., Pan, J. and Needleman, A. (1982), Void nucleation effects on shear localization in porous plastic
solids, Int. J. Fract., 19, 163.
Tvergaard, V. (1981), Influence of voids on shear band instabilities under plane strain conditions, Int. J. Fract.,
17, 389.
Tvergaard, V. (1982), On localization in ductile materials containing spherical voids, Int. J. Fract., 18, 237.
Tvergaard, V. and Needleman, A. (1984), Analysis of the cup-cone fracture in a round tensile bar, Acta Metal.,
32, 157.
Yamamoto, H. (1978), Conditions for shear localization in the ductile fracture of void-containing materials, Int.
J. Fract., 11, 347.
This page intentionally left blank.
A DILATATIONAL PLASTICITY THEORY FOR ALUMINUM
SHEETS

B. CHEN
Department of Mechanical and Industrial Engineering,
University of Illinois, Urbana, IL 61801, U.S.A
P. D. WU
Alcan International Ltd., P.O. Box 8400, Kingston, Ontario
K7L 5L9, Canada
Z. C. XIA
Ford Research Lab, P. O. Box 2053, MD 3135/SRL, Dearborn,
MI 48121, U.S.A
S. R. MAC EWEN
Alcan International Ltd., P.O. Box 8400, Kingston, Ontario
K7L 5L9, Canada
S. C. TANG
Ford Research Lab, P.O. Box 2053, MD 3135/SRL, Dearborn,
MI 48121, U.S.A
AND
Y. HUANG
Department of Mechanical and Industrial Engineering,
University of Illinois, Urbana, IL 61801, U.S.A

Abstract. The nucleation, growth and coalescence of micro-voids are im-


portant failure mechanisms in ductile materials. Gurson (1977) has devel-
oped a dilatation plasticity theory to quantitatively characterize the state
of deformation and damage associated with micro-voids in isotropic mate-
rials. This theory, however, is not applicable to aluminum sheets because
they are highly anisotropic. A dilatational plasticity theory for anisotropic
ductile materials is developed in this study. The constitutive law is es-
tablished for aluminum sheets that contain micro-voids, where the matrix
material of aluminum is characterized by an anisotropic constitutive model
developed by Barlat et al. (1991). Based on the numerical analysis, an ap-
proximate yield function is given in the closed form for anisotropic sheets.
17
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 1730.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
18 B. CHEN ET AL.

It shows that the mean hydrostatic stress plays an important role in the
plastic behavior of anisotropic micro-voided aluminum sheets.

1. Introduction

The ductile failure mechanism in most structural metals is the nucleation,


growth and coalescence of microvoids that result from debonding and crack-
ing of second-phase particles. It is well established that the growth of mi-
crovoids is governed by the hydrostatic stress field around the voids (e.g.,
Rice and Tracey, 1969; Huang, 1991). Gurson (1977) has incorporated this
effect of void growth in the isotropic J2 continuum plasticity theories, and
has developed a constitutive law for voided, dilating ductile materials. Un-
like the classical plasticity theories, the hydrostatic stress component, kk ,
as well as the void volume fraction, , has appeared in Gursons (1977)
dilatational plasticity theory. The Gursons theory has been widely used
in the study of ductile failure of solids due to void growth or plastic flow
localization (Needleman and Rice, 1978). The recent development on this
subject can be found in the review article by Tvergaard (1990).
There is an increasing need in recent years to use more aluminum al-
loys in industry because of their high strength/weight ratio. Similar to
other ductile materials, void nucleation, growth and coalescence also gov-
ern the ductile failure of aluminum alloys. The constitutive behavior of
aluminum, however, is highly anisotropic and cannot be characterized by
Hills quadratic (1950) or non-quadratic (1979) yield functions (e.g., Mel-
lor and Parmar, 1978; Mellor, 1981; Barlat et al., 1997). For example, the
theoretical forming limit strains predicted by isotropic yield functions are
unrealistically too high or too low. For this reason, Hosford (1979), Barlat
and co-workers (1989, 1991, 1997), Karafillis and Boyce (1993) have made
a series effort to develop the constitutive laws that are much more suitable
for aluminum alloys.
Gursons (1977) dilatational plasticity theory is not applicable to alu-
minum alloys because it is based on the isotropic J2 plasticity theories.
There are very limited studies on the Gurson-type dilatational plasticity
theory for aluminum because its anisotropic constitutive models are usu-
ally too complicated to provide analytical solutions for void growth. Liao
et al. (1997) have recently developed the approximate yield criteria for
anisotropic porous ductile sheet metals based on Hills quadratic (1950) as
well as non-quadratic yield functions (1979). The yield criteria are similar
to the Gursons (1977), except that the coefficient scaling the hydrostatic
stress is different to reflect the anisotropy.
PLASTICITY THEORY FOR ALUMINUM SHEETS 19

In order to determine the state of deformation and damage in aluminum


alloys, we develop a dilatational plasticity theory in this paper based on
Barlat and co-workers anisotropic plasticity theory. We begin with a sum-
mary of Barlat and co-workers plasticity theory in Section 2, followed by
the development of plane-stress anisotropic dilatational plasticity theory
for aluminum sheets in subsequent sections. The proposed dilatational plas-
ticity theory provides a means to quantitatively characterize the effect of
microvoid damage in aluminum alloys.

2. Anisotropic Plasticity Theories for Aluminum Sheets

Barlat and co-workers have developed the yield functions for anisotropic
aluminum alloys (Barlat and Lian, 1989; Barlat et al., 1991; 1997). For
example, Barlat and Lian (1989) and Barlat et al. (1991) have introduced
yield functions that account for the differences of yield strengths in the
rolling, transverse and normal directions. Some additional flexibility to the
yield function of Karafillis and Boyce (1993) was recently introduced in the
Barlat et al. (1997) model in order to account for the differences in the
yield stress under pure shear for two types of Al-Mg alloys that have the
same uniaxial yield stresses as well as the same equibiaxial yield stresses. As
an initial attempt to develop a Gurson-type dilatational plasticity theory
characterizing the effect of porosity on plastic yielding for anisotropic alu-
minum alloys, we use the 6-component anisotropic yield function proposed
by Barlat et al. (1991). Since the thickness of aluminum sheets is typically
much smaller than the in-plane dimension, we adopt the same assumption
as Liao et al. (1997) that the aluminum sheets are under the plane-stress
deformation. For simplicity, the plastic anisotropy within the sheet plane is
neglected, i.e., the aluminum sheets are modeled as transversely isotropic.
Let 11 , 22 , and 12 = 21 represent a plane-stress state ( 33 = 0) in
the sheet plane and x 3 be the out-of plane direction. In order to account
for anisotropy in plastic yielding, Barlat et al. (1991) have introduced a
transformed stress state as

(1)

where c 1, c 3 and c 6 are non-dimensional material constants that reflect the


anisotropy of the solid, and the in-plane isotropy requires At
20 B. CHEN ET AL.

the limit of c1 = 1 and c 3 = 1, the transformed stress state s degenerates


to the deviatoric stress and the material degenerates
to an isotropic solid. For aluminum sheets, the coefficient c3 can be more
than 20% higher than c 1 (Barlat et al. 1997).
The yield function is given in terms of the transformed stress state by
(Barlat et al., 1991)

(2)
where s1 , s2 and s3 are the principal values of the transformed state (s11 , s22,
s 33 , s12 ), i.e.,

(3)

Y is the uniaxial yield stress in the sheet plane (e.g., rolling direction),
and depends on the anisotropy constants c1 and c 3 and is given by

(4)

The elastic deformation is neglected such that the plastic strain rate
becomes the same as the total strain rate. The plastic strain rate is normal
to yield surface, and is given by

(5)

where , = 1, 2, and the proportionality coefficient represents the am-


plitude of plastic flow. The yield function (2) then gives the in-plane plastic
strain rate as

(6)
PLASTICITY THEORY FOR ALUMINUM SHEETS 21

where 12 = 2 12 is the engineering shear strain rate, and

(7)

The plastic work dissipation = can be evaluated via (6) to give


the following simple expression in terms of the non-dimentional pareameter
in(4), uniaxial yield stress Y in the sheet plane, and the proportionality
coefficient

(8)
The -value, defined as the ratio of transverse plastic strain-rate to the
through-thickness plastic strain-rate, is obtained in terms of anisotropy
constants c 1 and c 3 as

(9)

The theoretical limit strains predicted by the above constitution model


are in reasonable agreement with the experimental data for aluminum (Bar-
lat et al., 1991). Accordingly, the constitute law described above is used to
characterize the matrix material that contains microvoids in order to es-
tablish the Gurson-type dilatational plasticity theory for aluminum.

3. The Gurson-type Dilatational Plasticity for Aluminum Sheets


We extend Gursons (1977) approach to establish the approximate yield
function for an anisotropic aluminum sheet that contains microvoids. The
sheet is under the plane-stress deformation. A finite circular sheet contain-
ing a single through-thickness hole is subjected to a general macroscopic
strain rate on its outer boundary. The radius of the void is a, while
the outer radius of the matrix is b. Their ratio, a/b, is related to the void
volume fraction by

(10)
The microscopic strain rate is referred to the strain rate in the matrix
and is nonuniform due to the existence of the void. The matrix material
is characterized by Barlat et al.s (1991) anisotropic plasticity model. An
approximate, upper bound solution is obtained for the microscopic plastic
work dissipation Its integration over the matrix material gives the macro-
scopic work dissipation whose derivative with respect to yields the
stress state at the macoroscale.
22 B. CHEN ET AL.

The macroscopic strain rate can be decomposed to the volumetric


part and deviatoric part for plane-stress
deformation. The microscopic strain rate , which are nonuniform, can
be similarly decomposed to the volume-changing (constant shape) part
and the shape-changing (constant volume) part

( 1 1)

where such that


Following Gurson (1977) and Liao et al. (1997), we approximate the
shape-changing part at the microscale by the uniform field of macro-
scopic deviatoric strain rate

(12)

Also following Gurson (1977) and Liao et al. (1997), the volume-changing
part at the microscale corresponds to an axisymmetric velocity field,

(13)

which gives the strain rates

(14)

where r is the polar radius measured from the center of the void, A1 a n d A 2
are coefficients to be determined in the following. Matching of the imposed
deformation on the outer surface r = b of the matrix gives

(15)

where the right hand side represents the velocity associated with the
macroscopic volumetric strain field . The other relation between A 1
and A 2 is derived from the traction-free condition on the void surface. In
fact, the only non-zero stress component on the void surface is r =a ,
which gives the transformed stress state in (1) as ( s rr, s , s33 , s r ) =
The constitutive relation (6) then gives the
strain rates and on the void surface. Their ratio yields the
second relation between A 1 a nd A 2 as

(16)

where the strain field in (14) has been used, and


PLASTICITY THEORY FOR ALUMINUM SHEETS 23

(17)
is a non-dimensional parameter depending on the anisotropy constants c1
and c3 . The parameters A1 and A 2 are determined from (15) and (16) as

(18)

where is the void volume fraction. The microscopic strain rates,


accounting for both the volume-changing part and the shape-changing part,
are found in terms of the macroscopic strain rates as

(19)

The determination of microscopic plastic work dissipation is equiva-


lent to finding the proportionality coefficient , according to (7). It is ob-
served that, once are obtained as in (19), there remain four unknowns,
namely , s 11 , s 22 and s 12 , which are to be determined from three consti-
tutive relations in (6) and the yield function in (2). The eliminination of s 11,
s 22 a n d s 12 yields the following solution for

(20)

where

(21)

and g is an implicit function of strain rates , position (r, ), and the void
volume fraction , and is governed by the following nonlinear equation,

(22)
24 B. CHEN ET AL.

and

(23)

The macroscopic plastic work dissipation is the average of its micro-


scopic counterpart over the entire cell, i.e.,

(24)
where is given in (20).
The upper bound analyses of Gurson (1977) and Liao et al. (1997) are
used to calculate the macroscopic stress, , i.e.,

(25)

The integration with respect to r and in (25) must be evaluated numer-


ically. For any given macroscopic strain rates , (25) gives the corre-
sponding macroscopic stress state.
The macroscopic effective stress e is defined in the same way as the
microscale yield function in (2), i.e.

(26)

where is given in (4) and S1 , S 2 and S 3 are the principal values of the
transformed stresses S 11 , S 22 , S 33 a n d S 12 at the macroscale, which are
related to the macroscopic stresses via the linear transformation (1),

(27)
PLASTICITY THEORY FOR ALUMINUM SHEETS 25

The substitution of (27) into (26) gives the effective stress e in terms of
the stress components ,

(28)

where and For the


sheet subjected to the uniaxial tension 11 in the sheet plane, the effective
stress in (28) degenerates to the uniaxial stress 11 . It should be pointed
out that the effective stress depends on the anisotropy constants c1 and c 3
only through their ratio,
For each given macroscopic strain rate tensor , the macroscopic
effective stress e and the in-plane mean stress can only be obtained
numerically, and therefore yields an implicit relation between the two. This
relation reflects the dependence of the macroscopic yield function on the
mean stress and the void volume fraction for an anisotropic material and
is presented in the next section, along with an approximate but analytic
yield function.

4. An Approximate Yield Function and the Numerical Results


The relation between the macroscopic effective stress e a n d t h e m e a n
stress established in the previous section is fully implicit. It is de-
sirable to derive an approximate but analytical expression for the yield
function. Following Gurson (1977) and Liao et al. (1997), we derive an ap-
proximate relation between e and in this section and compare it
with the numerical results from the analysis in Section 3.
Consider the same geometrical model of a sheet containing a through-
thickness hole of the radius a as in Section 3. The outer radius of the
sheet is denoted by b. An axisymmetric velocity field, corresponding to
the macroscopic strain rate 11 = 22, is imposed on the outer boundary
r = b. The sheet is also subjected to a uniform strain rate 33 in the out-of-
plane direction. It should be pointed out the deformation is not plane-stress
anymore because of the imposed 33 . The microscopic strain field in the
matrix is then given by

(29)

where the coefficient A'1 is determined by the incompressibility of the plastic


deformation,

(30)
26 B. CHEN ET AL.

The other coefficient, A' 2, is determined by the imposed velocity field on


the outer boundary r = b,

(31)

which is similar to (15).


The rest of the analysis is almost identical to that in Section 3, except
that it is not plane stress ( 33 0) but axisymmetric. The matrix material
is characterized by the Barlat et al. (1991) model, and the non-vanishing
stresses on the microscale are rr, and 33 . The corresponding non-
vanishsing macroscopic stresses are 11 = 22 and 33. Therefore, the
in-plane mean stress is while the effective stress defined by
(26) becomes

(32)

Unlike Gurson (1977) and Liao et als (1997) analyses which yield the
closed-form solutions for the approximate yield functions, the present ax-
isymmetric analysis still does not warrant analytical solutions because the
yield functions in (2) are highly nonlinear. However, the analytical solutions
can be found for the following two limits:
(1) uniaxial tension, It is straightforward to show
A' 2 = 0 from (30) and (31) such that the microscale is also subjected to
uniaxial tension, This makes sense because the existence
of the through-thickness hole does not violate the condition rr = = 0
in uniaxial tension along the thickness direction. The macroscopic in-plane
stresses, 11 and 22, are the average of their microscale counterparts, 11
and 22 , within the unit cell (Gurson, 1977), and therefore also vanish. The
closed-form solution gives the macroscopic effective stress as e = (1) , Y
which also makes sense since the factor 1 accounts for the reduced area
to sustain the load. At this limit, the in-plane mean stress and
the effective stress e = (1 ) Y .
(2) plane-strain deformation, 33 = 0; It is evident from (29) the microscale
deformation is also under the plane-strain condition, 33 = 0. In conjuction
with the normality of plastic flow, 33 = 0 gives that the microscale stress
in the thickness direction is exactly the same as in-plane mean stress, i.e.,
By averaging over the unit cell, this condition
becomes which gives a vanishing effective stress, e = 0
from (32). The closed-form solution also gives the in-plane mean stress as
where is given in (4) in terms of the anisotropy
constants c1 a n d c3.
PLASTICITY THEORY FOR ALUMINUM SHEETS 27

These two limits lead to a very natural construction of the approximate


yield criteria for the entire range of in-plane mean stress and effec-
tive stress e. It is observed that the first limit is equivalent to
and The second limit, after some manipulations, can
be written as and The ap-
proximate yield criterion that satisfies these tow limits and bears similarity
with Gursons (1977) and Liao et als (1997) yield criteria for dilatational
plasticity is simply the combination of the above two limits,

(33)

In particular, this relation is exact at the above two limits, i.e., e =


and In fact, (33)
is a rather accurate representation of the numerical solutions for all com-
binations of strain rates. The yield function in (33) is extremely similar to
those established by Gurson (1977) and by Liao et al. (1997), though the
effective stress is defined by the Barlat et al. model (1991) via (28). The
anisotropy comes into play through the coefficient to scale the
in-plane mean stress as well as through the effective stress in (28).
It is also observed that the yield function in (33) depends on the ratio of
anisotropy constants c 1 a n d c 3 .
It is recalled that the aim of the present study is to establish the yield
function for plane-stress aluminum sheets containing through-thickness
holes. Therefore, it is important to compare the above approximate yield
function established from the axisymmetric analysis with the numerical so-
lutions for plane-stress aluminum sheets. The relation between the effective
stress e and the in-plane mean stress from the previous section for
the plane-stress is shown in Fig. 1. The approximate yield function in (33)
is also presented in Fig. 1 for comparison. Both the effective stress and the
in-plane mean stress are normalized by the uniaxial yield stress Y i n t h e
sheet plane. The void volume fraction has a rather large variation, from
0.02 to 0.2, while the anisotropy coefficients are c 1 = 1 and c 3 = 1.2, which
gives a very large -value = 1.80 from (9).
It is clearly observed that, without any parameter fitting, the numerical
results for plane-stress sheets agree very well with the approximate yield
function in (33) for a large range of void volume fraction . We have also
examined other combinations of the anisotropy constants, The numeri-
cal results all confirm that the yield function is well approximated by (33).
Therefore, (33) gives a rather accurate measure of the yield function for
anisotropic aluminum sheets.
28 B. CHEN ET AL.

Figure 1. The yield function for anisotropic aluminum sheets containing microvoids; the
matrix material is characterized by Barlat et al.s (1991) anisotropic constitutive model;
e a n d are the effective stress and in-plane mean stress, respectively; Y is the
uniaxial yield stress in the sheet plane; is the void volume fraction; the anisotropy
c o n s t a n t s a r e c1 = 1 and c 3 = 1.2, p is given in (4), and the value is 1.80 from (9).

It should be pointed out that the numerical results presented in Fig. 1 for
plane-stress sheets have covered all possible combinations of macroscopic
strain rates, ranging from with 12 = 0. In fact the left

limit corresponds to while the right limit


corresponds to the most right point on each curve.
The yield function (33) for anisotropic sheets has a remarkable resem-
blance to that of Gurson (1977) for isotropic solids, even though the effective
stress is defined by the Barlat et al. (1991) model via (28) instead of the
Von Mises effective stress. In fact, its interception with the vertical ( e)
axis, 1 , is identical to that in the Gurson model. The interception with
the horizontal axis also has the same dependence on the void vol-
ume fraction, In 1 ; though the associated coefficient is slightly different,
reflecting the effect of the plastic anisotropy. The curves between these two
limits are both characterized by the cosh function in (33) and in the Gurson
PLASTICITY THEORY FOR ALUMINUM SHEETS 29

model. Similar conclusions have been established by Liao et al. (1997) for
Hills quadratic (1950) and non-quadratic (1979) anisotropic yield criteria.

5. Conclusion

We have generalized Gursons (1977) isotropic dilatational plasticity theory


to the anisotropic aluminum sheets in this paper. Following the same ap-
proach of Gurson (1977) and Liao et al. (1997), we have analyzed a matrix
containing a single void subjected to an imposed deformation on the outer
boundary. The anisotropic constitutive model of Barlat et al. (1991) is used
to characterize the plastic behavior of aluminum matrix. The relation be-
tween the effective stress e and the in-plane mean stress is fully
implicit and must be obtained numerically. However, we have obtained an
approximate yield function that is rather accurate for the entire range of e
and This approximate yield function is similar to that in Gursons
model, but the effect of plastic anisotropy has been accounted for.

Acknowledgements

Y.H. gratefully acknowledges the research grants from Ford Foundation and
from Alcan Int. Ltd.

References
Barlat, F. and Lian, J. (1989) Plastic behavior and stretchability of sheet metals. Part I:
A yield function for orthotropic sheets under plane stress conditions, Int. J. Plasticity
5, 51-66.
Barlat, F., Lege, D.J., and Brem, J.C. (1991) A six-component yield function for
anisotropic materials , Int. J. Plast 7 , 693-712.
Barlat, F., Maeda, Y., Chung, K., Yanagawa, M., Brem, J. C., Hayashida, Y., Lege,
D.J., Matsui, K., Murtha, S. J., Hattori, S., Becker, R.C. and Makosey, S. (1997)
Yield function development for aluminum alloy sheets, J. Mech. Phys. Solids 4 5 , No.
11/12, 1727-1763.
Gurson, A.L. (1977) Continuum theory of ductile rupture by void nucleation and growth:
part I - yield criteria and flow rules for porous ductile media, J. Eng. Mater. Tech
99, 2-15.
Hill, R. (1950) The mathematical theory of plasticity, Oxford University Press , L o n d o n .
Hill, R. (1979) Theoretical plasticity of textured aggregates, Math. Proc. Camb. Philos.
Soc. 8 5, pp. 179.
Hosford, W.F. (1979) On yield loci of anisotropic cubic metals, Proc. 7th North American
Metalworking Conf., SME, Dearborn, MI, pp. 191-197.
Huang, Y. (1991) Accurate dilatation rate for spherical voids in triaxial stress fields, J .
Appl. Mech. 5 8 , 1084-1086.
Karafillis, A. P. and Boyce, M. C. (1993) A general anisotropic yield criterion using
bounds and a transformation weighting tensor, J. Mech. Phys. Solids 4 1, 1859-1886.
Liao, K.-C., Pan, J., Tang, S.C. (1997) Approximate yield criteria for anisotropic porous
ductile sheet metals, Mechanics of Materials 2 6 , 213-226.
Mellor, P.B. (1981) Sheet metal forming, Int. Metals Rev. 2 6 , 1-20.
30 B. CHEN ET AL.

Mellor, P. B. and Parmar, A. (1978) Plasticity of sheet metal forming, in D. P. Koistinen


and N. M. Wang (eds.), Mechanics of Sheet Metal Forming , Plenum Press, New York,
pp. 53-74.
Needleman, A. and Rice, J. R. (1978) Limits on ductility set by plastic flow localization,
in D. P. Koistinen and N.-M. Wang (eds.), Mechanics and Sheet Metal Forming 17 ,
Plenum, New York, pp. 237.
Rice, J.R. and Tracey, D.M. (1969) On the ductile enlargement of holes in triaxial stress
fields, J. Mech. Phys. Solids 1 7, 201-217.
Tvergaard, V. (1990) Material failure by void growth to coalescence, In: J. W. Hutchinson
and T.Y. Wu (eds.), Advances in Applied Mechanics 2 7, pp. 83.
INTERNAL HYDROGEN-INDUCED EMBRITTLEMENT IN IRON
SINGLE CRYSTALS

JIAN-SHENG WANG
Northwestern University
Evanston, IL 60201, USA

Abstract: A thermodynamic model for internal hydrogen-induced embrittlement (HIE)


in single crystals is proposed. The model is based on the assumption that the ductile
versus brittle transition is controlled by the competition between dislocation emission
from the crack tip and cleavage decohesion of the lattice. Embrittlement in single
crystals is induced by segregation of hydrogen in solid solution to the crack tip and/or
the fracture surfaces during separation, which reduces the cohesive energy of the lattice.
This process will occur when the mobility of hydrogen atoms is high so that a surface
excess of hydrogen can be built up during separation. The model predictions for
hydrogen induced cleavage in iron single crystals are presented.

1. Introduction

Hydrogen-induced embrittlement (HIE) has been observed not only in polycrystalline


metals and alloys, but also in single crystals, e.g., in single crystals of Ni [1, 2], Ni-
based alloys [3, 4], Fe and FeSi alloys [1, 5-9], stainless steels [10], and intermetallics
[11]. For a system where the formation of a hydride is thermodynamically unattainable
or kinetically impractical, solution hydrogen-induced embrittlement may occur due to
precipitation of gaseous hydrogen or methane; localized plastic deformation prompted
by the interaction between hydrogen atoms and dislocations at the crack tip; or due to
the reduction of the lattice or grain boundary cohesion. The hydrogen-enhanced
localized plasticity theory suggests that hydrogen in a solid solution reduces the barrier
to dislocation motion through an elastic shielding effect [12-14], thereby increasing the
amount of plastic deformation that occurs in a localized region adjacent to the fracture
surface, causing embrittlement. In contrast, the cohesion-reduction theory postulates
that segregation or adsorption of hydrogen decreases the cohesive energy inducing
embrittlement [e.g. 15-18].
Vehoff pointed out, in a recent review, that for HIE to occur, hydrogen has to
enter the fracture processing zone (FPZ) to reduce the local atomic bonding strength at
the crack tip [18]. In the presence of solute hydrogen in polycrystals, this can occur due
to segregation of hydrogen to grain boundaries [15, 16]. In the presence of external
gaseous hydrogen, this can occur due to adsorption of monatomic hydrogen at the
newly formed fracture surfaces and the FPZ [2]. While the cohesion-reduction theory
can explain hydrogen induced cleavage-like fracture in single crystals in the presence of
external hydrogen gas, a rigorous thermodynamic analysis of the cohesion-reduction
31
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 3147.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
32 J.-S. WANG

theory for HIE in single crystals in the absence of external hydrogen gas is not
available. The present work suggests a thermodynamic model for HIE in single crystals
when hydrogen is present in the solid solution and any damage induced by hydrogen
charging, or any softening-hardening effect is negligible. This model is a natural
extension of the thermodynamic analysis of segregation-induced interfacial
embrittlement of Rice, to whom this book is dedicated, and his colleagues [15, 19-22].
To understand how a single crystal can be embrittled by hydrogen in solid solution, it
might be worthwhile to review how Rice resolved the dilemma of segregation-induced
interfacial embrittlement.

2. Thermodynamics of Segregation-Induced Interfacial Embrittlement

How grain boundary segregation could induce embrittlement was a puzzle less than
three decades ago. Thermodynamics asserts that a lower energy state is more stable. For
example, special grain boundaries, such as 3, 11 coincide lattice site (CLS) grain
boundaries, are more stable than random grain boundaries because of their low grain
boundary energies. Segregation reduces the grain boundary energy; it would, intuitively,
stabilize the grain boundary, but why, instead, does it promote intergranular brittle
fracture? The cohesive energy of a grain boundary, or equivalently, the reversible work
of intergranular separation, is conventionally (but, as Rice pointed out, not completely)
defined as
c = 2 s b (1)

where s and b are the surface and grain boundary energies, respectively. For a low
temperature or fast interfacial separation process, when redistribution of the segregant at
the newly created fracture surfaces is unattainable because of its low mobility, s would
remain unchanged. In this case, a reduction in b by segregation would increase c.
Based on an incomplete thermodynamic analysis, it has been claimed that low
temperature work of grain boundary fracture is independent of segregation [23],
contradictory to experimental observations. A rigorous thermodynamic analysis of
interfacial cohesion in the presence of solute atoms was provided by Rice and through
this, the problem of segregation induced intergranular fracture was solved [19].
In his brilliant treatment of the interfacial
cohesion problem, Rice introduced two new
thermodynamic variables: the stress acting at the
interface and the separation distance of the
creating surfaces. The force per unit area acting
on the separating atoms at the interface, , is a
function of the separation distance . T h e
cohesive energy of the interface, c , is defined
by the reversible work of separation, i.e., the
area under the curve:

(2)
Fig. 1. Schematic of the
relation at a crack-tip.
where 0 is the initial separation of the unstressed interface and is the separation
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 33

excess under the stress (Fig. 1). Rice pointed out that the most important parameters
associated with crack initiation were the cohesive strength of the interface c , i.e., the
maximum of the curve, and the cohesive energy of the interface, c . Since the
relation is affected by segregation, both c and c may alter in the presence of
segregants.
Interface cohesion is not a state function in general, but depends on the
thermodynamic path followed by separation. There are two limiting cases in interfacial
separation processes in the presence of segregants: separation at constant interface
concentration and separation at constant chemical potential. These two limiting cases
identify two different thermodynamic paths. The first path is a fast separation on a
time scale which does not allow further matter transport to the interface (the immobile
case). The second is a slow separation on a time scale which allows full composition
equilibrium between the interface and a matter source at a constant potential (the
mobile case). By introducing two new valuables, and , Rice derived that

(3)

for separation at constant , where 0 ( ) is the potential corresponding to excess


concentration on the unstressed interface and ( ) the potential corresponding to the
net excess concentration on the two completely separated surfaces, and

(4)

for separation at constant , where () is the initial segregant excess on the unstressed
interface and () is the equilibrium excess on the two completely separated free
surfaces.
Rices analytic results were later given in terms of a reversible work cycle in
chemical potential-composition space by Hirth and Rice [20] as shown in Fig. 2. The
rigorous thermodynamic analysis of Hirth and Rice demonstrated that for a fast
separation at constant , the change in the cohesive energy is

(5)

corresponding to area OAYO along the = b( ) curve in Fig. 2. For a slow


separation at fixed the change in the cohesive energy is
(6)

corresponding to area OBYO along the = s (/2) curve in Fig. 2. Here s () gives the
segregant excess on a single free surface, and b () on a grain boundary, at
equilibrating potential .
34 J.-S. WANG

Equations (5) and (6) link c , for separation at fixed composition or at fixed chemical
potential, to quantities, which can, in principle, be estimated from solute segregation
studies. Normally the potential necessary to equilibrate on a grain boundary will be
larger than the potential to equilibrate the same amount on a pair of free surfaces (at /2
on each), i.e., b ( ) > s ( /2), and also 2s () > b (). With this normal type of
segregation behavior c > 0 for both paths, the segregation reduces c , and thus is
expected to promote embrittlement.
The analysis of Hirth and Rice shows that for an interface with solute segregation,
the definition for grain boundary cohesion given in (1) is valid only for the mobile
case of fully equilibrated separation at constant chemical potential. Here, s and b
corresponds to the free energy of the completely separated surfaces and the unstressed
interface, respectively, each of which is in equilibrium with the potential source, i.e.,

(7)

Since a normal segregant reduces the free energy of the surface more than reducing the
free energy of the grain boundary, the cohesive energy of the interface is thus reduced.
Equation (1) is, however, invalid for fast separation when further matter transport is
not allowed during separation. In this immobile and non-equilibrium case, the
cohesive energy of the interface is given by

(8)

where b ( 0 ) is the free energy of the interface in which solute of concentration 0


equilibrates with the bulk phase, s ( 0 /2) is the free energy of the surface in which
solute of concentration 0 /2 equilibrates with the surface but not the bulk phase, and
s and b are corresponding chemical potentials. Since s (0 /2) < s0 , where s0 is the
surface free energy without segregation, and for a normal segregant, the last term in the
right hand side of (8) is negative, interface cohesion is reduced. The early dilemma in
understanding segregation induced intergranular fracture is thus resolved.

Fig. 2. Schematic of the potential-excess spaces for a grain boundary and free
surfaces. The original state is O( b , 0 ). Two limiting cases and a transient
case for grain boundary separation are shown.
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 35

The more general case is that both the grain boundary excess and the chemical potential
vary during separation following a transient path. Referring to Fig. 2, the separation
process is a transition starting from the initial state O( b , 0 ) to a non-equilibrium state
T( t , t ). The change in the cohesive energy of the interface corresponds to area OTYO,
where t ( t ) is the transient chemical potential of the segregant at a non-equilibrium
surface excess of t . Based on segregation kinetics, a model has been developed to
evaluate the embrittlement propensity under transient conditions [22].

3. Thermodynamics of Segregation-Induced Embrittlement in Single Crystals

Considering now a single crystal with a segregant in solid solution, a crack is initiated
and propagating under stress leading to fracture, the temperature is high or the
separation is slow under the mobile condition so that chemical potential is constant
during matter transport. Following the same procedure of Hirth and Rice, the
thermodynamics for segregation-induced embrittlement in single crystals is described as
following.
Thermodynamics states that for reversible alteration of state of a system

dU = TdS + dw rev (9)

where dw rev is the reversible work with the sign opposite to the conventional chemical
thermodynamics usage. In an isothermal system consisting of a single crystal m, and a
segregant of chemical potential , and capable of changing the surface area, dAs , under
the surface tension, s , in the absence of any external device work

where V is the volume of the system, P is the uniform pressure, dn and dn m are the
exchanges of matter in moles. The surface tension is identical to the surface free energy
if any change in the surface area by elastic stretching bonds can be ignored and,
henceforth, we call s the surface free energy. Equation (9) becomes

and in terms of the Helmholtz free energy, F = UTS

The extensive quantities F, V and n can be divided into bulk quantities and surface
excess quantities normalized to unit area of the surface, f, v, and and m , thus

(10)

A simple argument of Hirth and Rice demonstrated that integration of (10) at constant T
yields
(11)
36 J.-S. WANG

Differentiation of (11) and by making use of the Gibbs-Duhen relation, it is easy to


show that at constant T

(12)

Considering that the dividing surface can be chosen so that v = 0 (in the Gibbs sense) or
the term vdP can be disregarded (in the Guggenheim sense), and that the phase law does
not allow d m independent of d, (12) reduces to

(13)

Integration of (13) leads to the surface free energy

(14)

where s0 is the surface free energy of the pure single crystal and 0 RT lnc 0 is the
potential of the solute of concentration c 0 in the bulk. The validity of this relation to
fracture of a single crystal relies on the mobile condition that the chemical potential
remains constant. Under this condition, the change in the cohesive energy of the lattice
is

(15)

where = s () is the excess on a single surface. The separation processes start from
the origin O at constant = 0 and end at B with surface excess s = . The change in
cohesive energy is represented by area OBYO along the = s ( ) curve in the chemical
potential-excess space (Fig. 3).

The original analysis of Hirth and Rice for a system with an initial interface (a grain

Fig. 3. Schematic of the potential-excess space for a free surface. Cleavage


fracture at constant chemical potential or under transient conditions is shown.
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 37

boundary) is valid for a system without a grain boundary, because thermodynamics does
not distinguish the physical nature of the interface: no matter if it is a grain boundary or
a surface. For a system without an initial grain boundary, (15) can be derived simply by
setting the b term in (6) to zero. The same is true for Rices treatment with and as
variables. Similarly, by setting = 0 in (3) and (4), the cohesion of a single crystal
under immobile and mobile conditions can be derived. Under immobile
conditions, dc /d in (3) is identically zero, the cohesive energy of the crystal is
independent of segregation if separation is fast. It ought to be understood that this
statement is true if any changes in dislocation behavior, the bonding nature and the
lattice distortion et cetera, induced by solute atoms can be ignored.

4. Micromechanics of HIE in Fe Single Crystals

A micromechanical description of the ductile versus brittle fracture of a crystal concerns


dislocation emission from the crack tip. It is generally believed that the tip response of a
stressed crack is governed by the competition between dislocation nucleation from the
tip and cleavage decohesion. This concept was modeled by Rice and Thomson [24] and
advanced by Schoeck [25] and Rice [26]. In general, the critical energy release rate for
dislocation nucleation from the crack tip, G disl , and the critical energy release rate for
Griffith cleavage decohesion of the crystal, Gc l e a v , is compared. If Gdisl < G c l e a v ,
dislocation emission is predicted to occur first as the crack tip loading is increased,
thereby blunting the crack tip and reducing the tip stress field required for cleavage. In
this case, the crystal is interpreted as intrinsically ductile and the crack tip is a
dislocation emitting tip. Alternately, if G disl > G cleav , atomic decohesion occurs first,
producing cleavage, or in essence, cleavage is potentially conceivable. In this case, the
crystal is esteemed as intrinsically brittle and the crack tip is called a non-emitting tip.
Understanding the HIE is thus reduced to evaluating how Gdisl and G c l e a v are affected by
segregation of hydrogen to the crack tip and the fracture surfaces.

4.1. THE SEGREGATION INDUCED REDUCTION OF G c l e a v

It has been derived from the cohesive zone model that the energy release rate for
cleavage decohesion is equivalent to the cohesive energy given by (2), providing that
self atom trapping at the crack tip is negligible, i.e.,

G cleav = c (16)

Because of the reduction of c , ductile-brittle transition may occur if the inequality of


G disl < Gcleav for ductile behavior is reversed to Gdisl > Gc l e a v for brittle behavior.
When the surface excess is less than values corresponding to full occupancy of a set
of segregation sites, idealized as all having the same low energy relative to solute sites
in the bulk, the simple Langmuir-McLean model [27, 28] may apply,

An intrinsically ductile solid is not necessarily to fracture in a ductile manner. If the dislocation mobility is
low, the newly formed dislocation may not be able to move away from the tip, exerting an image force to the
tip and preventing further nucleation of dislocations from the tip, resulting in a brittle behavior.
38 J.-S. WANG

(17)

where is the surface saturation excess and the inherently negative segregation free
0

energy g s is referenced to a bulk phase at the same T, i.e., is based on the expression
= RT ln [x/( 1x)] RT ln (x) for the equilibrating potential when a fraction x of
available solute sites is occupied in the bulk and x<<1. Integration of (15) yields

(18)

for an equilibrium separation, where s / 0 is the surface coverage:

(19)

It has been derived [22] that for a grain boundary separating at a constant chemical
potential = b ( b )

(20)

where is the equilibrium coverage in the grain boundary. Noting that 0 =


by setting b = 0 equation (20) for an interface reduces to (18) for a single crystal
lattice where no initial interface coverage exists.

Equation (18) gives the maximum reduction in cohesion when equilibrium segregation
occurs during separation. This may not be attainable when separation is fast or the
temperature is low. In reality, the more general case is that during separation the
chemical potential varies following a transient path OT in Fig. 3. A non-equilibrium
surface excess, t , is attained, which is controlled by segregation kinetics. Referring to
Fig. 3, the separation is a transition starting from the initial state O(c 0 , 0 ) along a non-
equilibrium path to a transient state T( t , t ). The change in the cohesive energy of the
lattice corresponds to area OTYO. To determine the reduction in cohesion under
transient conditions, t, a knowledge of the actual details of the trajectory OT, i.e., the
function = t ( ), is needed, which are governed by the diffusional transport of the
segregant from solid solution to the separating surfaces. In the case where no second
phase is involved during separation, the function t ( ) must be continuous, single
valued. Assuming a linear relation as a first order approximation, one may derive the
reduction of cohesion under transient conditions:

(21)

where t t / 0 is the transient coverage and s is the maximum attainable surface


coverage given by (19). The transient coverage might be derived from McLean
segregation kinetics:
(22)
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 39

with

where D is the diffusion coefficient of the segregant, the fracture process time, h the
thickness of the surface region and s /c 0 is the maximum surface enrichment factor.
The fracture process time, , is related to the crack growth rate, , by = , where
is the size of the cohesive zone at the crack tip. For an atomically sharp cleavage
crack, 3~4 lattice spacings [21].

4.2. EFFECTS OF SEGREGATION OF HYDROGEN ON G cleav IN IRON SINGLE


CRYSTALS

Hydrogen tends to occupy high symmetry


adsorption sites: three-fold coordinated sites
on (110) surfaces and bridge sites on (100)
surfaces. The hydrogen covered (100) surface
exhibits a c(2x2) reconstruction with a
saturation occupancy of 1:1 atomic ratio [29].
Suppose hydrogen is induced into iron single
crystals by cathodic charging at 298 K and
fugacity f = 300 bar, the potential-excess
space for cleaving along {100} plans is
shown in Fig. 4. The maximum reduction in
cohesion for the cleavage plane, calculated
by (19) and represented by the shadowed area
in Fig. 4, is 1.37 J/m. For comparison, the
b - space for a random grain boundary is
also shown in the figure. The maximum
reduction in cohesion for the grain boundary, Fig. 4. The potential-excess space for
calculated by (20), is 0.804 J/m. Under the Fe(100) plane. Hydrogen pre-charging:
T= 298 K and fugacity f =300 bar.
same conditions, the propensity for HIE in
single crystals might be greater than that in
polycrystalline specimens.
Under transient conditions, the reduction in cohesion is governed by the kinetics of
segregation. Therefore, it is a function of the crack propagation rate and temperature.
Figure 5 shows the reduction in cohesion of iron crystals versus the cracking rate at 298
and 373 K, where the cathodic charging is conducted during separation. Suppose a 5%
reduction in cohesion is necessary for ductile-to-brittle transition, corresponding to c
= 0.1826 J/m, then HIE in iron single crystals occurs at room temperature when the
4
crack propagation rate is quite low: 410 /sec, showing a time delayed cracking
behavior. Figure 6 shows the temperature dependency of the reduction in cohesion at a
fixed cracking rate = 0.1 /sec for crystals cathodically charged at 343 K. The
embrittlement intensity increases with increasing temperature because of the increase in
hydrogen diffusivity. It decreases with further increasing temperature because of
reduced segregation. The model predictions are consistent with generally observed
40 J.-S. WANG

experimental results. For comparison, the temperature dependency for polycrystals is


also plotted in the figure, where the specimen is cathodically charged under the same
conditions and equilibrium segregation of hydrogen in grain boundaries is assumed. For
the transient case, the embrittlement intensity for polycrystalline specimens is greater
than that for single crystals.

4.3. EFFECT OF SEGREGATION OF HYDROGEN ON G disl

The critical energy release rate for dislocation nucleation from a crack tip, Gdisl , could
be evaluated by the Rice-Thomson model. The original approach of Rice and Thomson
takes into account the balance of the work done by the applied stress and the energy of a
dislocation loop emanating from the crack tip. This procedure evokes a dislocation core
cutoff, a poorly defined parameter in the continuum elastic dislocation theory and, in
essence, it deals only with emission of dislocations from the crack tip, not dislocation
nucleation. The newly developed Peierls-Nabarro type approach [25, 26, 30-34] realizes
that a full dislocation is likely to emerge unstably from an incomplete, incipient

Fig. 5. The cracking rate dependency of HIE Fig. 6. The temperature dependency of HIE in
for Fe single crystals. Fe single crystals and polycrystalline specimens.

dislocation at the tip, and the barrier for nucleation of this incipient dislocation is
proportional to Peierls barrier for dislocation motion. The analytic and numerical
treatments of the model [26, 30] solve the elasticity problem of a traction free crack
with a Peierls stress versus displacement relation being satisfied as a boundary
condition along a slip plane ahead of a crack tip. Once this problem is solved for a
suitable constitutive relation for material sliding and perhaps opening along a slip plane,
there is no need for the core cut-off parameter. The Peierls-Nabarro type approach
describes that the critical energy release rate for dislocation initiation from the tip, Gdisl ,
is scaled by the unstable stacking energy, us , a newly introduced solid state physics
parameter [26]. Considering the softening effect of the mode I loading, atomistic
simulations by the Embedded Atom Method (EAM) for Fe derived the fully relaxed
unstable stacking energy us = 0.44 J/m [35].
Depending on the orientation of the active slip systems (the inclination angle
of the slip plane against the crack plane and the angle between the Burgers vector and
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 41

the crack front normal), Gdisl as a function of us can be obtained numerically [26, 31,
32, 35]. Suppose an atomically sharp crack at a (001) plane in an -Fe crystal with the

crack front parallel to the [110] direction and the crack propagates in the [110]
direction, the critical energy release rate for dislocation nucleation from the crack tip at
0 K was solved numerically: G disl = 3.55 J/m 2 [35]. Comparing with the cohesive energy
of the cleavage plane, c = 3.652 J/m 2 [36], pure iron is predicted to be intrinsically
ductile.
The lowest ductile versus brittle transition temperature (DBTT) ever measured
for -Fe is 163 K [37]. It has been recognized that the DBTT of iron is strongly
influenced by trace impurities, such as carbon, nitrogen, oxygen, sulfur, phosphorus,
boron, and hydrogen. When iron is sufficiently purified, the physical and chemical
properties are considerably different from accepted values. The experimentally
observed low temperature brittleness of iron is related to the inevitable impurities and
the low mobility of dislocations. Assuming a constant value of us ,for an ideally pure
iron single crystal at 0 K, the ductile versus brittle transition might occur if the cohesive
energy is reduced about 2.7%. At non-zero temperatures, Gdisl is lowered by thermal
activation, a further reduction in cohesion is needed for the transition. This behavior is
altered due to the presence of inevitable trace elements. Impurities affect not only the
cohesion, but also Gdisl though the influence of solute atoms on us and the dislocation
line energy. Some of the trace impurities may increase us and some of them may
decrease it. In addition, the pinning effect of large-atomic-size impurities on
dislocations may impede dislocation nucleation from the crack tip [38]. Quantitative
evaluation of the net effect of trace elements on the ductile versus brittle transition in Fe
single crystals requires comprehensive atomistic simulations, which have not been
available and is beyond the scope of the present work.
The effects of interstitial atoms on dislocation emission from a crack tip was
revealed by an atomic model in the spirit of the Peierls approach [39]. The model is
based on an EAM-type potential for nickel. The energetics of dislocation emission from
a crack-tip in nickel containing hydrogen is analyzed. The results show a substantial
effect on the unstable stacking energy as the dislocation passes an interstitial on the slip
plane, but the effect of an interstitial on the resistance to dislocation emission expressed
in terms of the maximum lattice resistance, r , is small and then only if it is confined to
a region very near the crack tip. Similar simulations for iron have not been available. It
is believed that the Peierls potential for screw dislocations in Fe single crystal is
reduced by hydrogen [40], indicating a reduction in us .
A thermodynamic model was proposed to evaluate the change in the
dislocation line energy due to segregation of solute, in particularly, segregation of
hydrogen [15]. Analogous to the surface or interface segregation, the reversible work of
forming a unit length of dislocation in the presence of a solute is given

(23)
with e0 equal to the work of formation in the absence of solute, d is the solute excess at
the dislocation and is the equilibrium chemical potential. Equation (23) is evaluated
for conditions where the dislocation is formed at chemical equilibrium between the
solute and the bulk and dislocation core sites. The McLean isotherm is used to define
the equilibrium excess d at the dislocation in terms of the lattice concentration, the
42 J.-S. WANG

segregation free energy g d to the dislocation sites, and the saturation excess level d0 at
the dislocation. The reduction in the line energy is evaluated as
(24)

where d = d / d , is the dislocation line coverage of the solute atoms.


0

For conditions where the dislocation core acts as segregation sites (i.e., g d, <
0), the effect of adding an impurity to the bulk material is seen to reduce Gdisl . This
effect is intended to complement impurity effects on atomic decohesion discussed
earlier, and on diffuse plastic flow from external, non-crack-tip dislocations. A
quantitative estimate of the effect of H on dislocation emission in Fe is difficult, since
0
d is poorly known. For a reference, the binding energy of H to a non-screw dislocation
in Fe is estimated as 58.6 kJ/mol [41], which is lower than the estimate for grain
boundaries but considerably higher than the estimate for free surfaces [22]. One could
expect that the hydrogen-induced reduction in Gdisl is less severe than in G cleav. The net
effect of hydrogen in iron single crystals is to reverse the inequality of Gdisl < G cleav for
an otherwise pure iron crystal to the inequality of G disl > Gcleav , or is to increase the
ductile versus brittle transition temperature for an impurity-containing iron crystal.

5. Discussion

The present thermodynamic and micromechanical model for HIE in single crystals
states that because of the segregation-induced reduction of cohesion, the condition for
an intrinsically emitting crack, Gdisl < Gcleav is reversed to G disl > G cleav for a non-
emitting crack, leading to brittle fracture or an increase in the DBTT. A tactic
assumption behind this statement is that while brittle fracture occurs, the crack tip is
elastic, despite the fact that pre-existing dislocations in the near-tip region may induce a
large amount of plastic deformation and hence, the brittle fracture energy of a metal is
usually orders of magnitude higher than the cohesive energy.
It has been realized that for a non-emitting crack, due to the microscopic
discreteness of plastic flow at a length scale too small for continuum plasticity, there
may exist an elastic enclave free of dislocations around the tip [42-45]. Because of the
shielding effect of pre-existing dislocations, when the local tensile stress at the tip is
great enough to meet conditions for Griffith cleavage, the corresponding concentrated
stress field near the tip contains large enough shear stresses to move pre-existing
dislocations. In other words, while conditions for cleavage are satisfied at the tip, a
plastic zone develops near the tip. The motion of those pre-existing dislocations induces
a strong shielding effect of the crack tip from the full effect of the externally applied
load. Therefore, the external load level for cleavage is greatly increased, i.e. Kapp > > Ktip
or equivalently, G far >> G tip . The shielding ratio, Gfar /G tip , is a strong function of c . A
small change in c causes a large change in the shielding ratio, i.e., the cohesive energy
serves as a control valve for fracture.
The self-consistent elastic enclave model provides a qualitatively sensible
description of cleavage in intrinsically cleavable materials in the presence of a large
amount of plastic deformation and insures the tip response upon stresses being governed
by the competition between dislocation emission and cleavage decohesion, provided
that the mobility of dislocations is not the limiting step. For an emitting crack tip, where
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 43

G disl < G cleav , such an elastic core does not exist 2 ; the conditions for Griffith cleavage
cannot be satisfied, resulting in a ductile fracture.
Interaction between hydrogen atoms and the near tip stress field has been a
great concern in the study of hydrogen assisted cracking. The near tip stress field may
increase the solubility of hydrogen in solid solution; enhance the local concentration of
hydrogen; accelerate hydrogen diffusion; promote segregation and therefore intensify
HIE. While a general view of stress-enhanced diffusion appears to be adverse [46-48], it
has been confirmed that the concentrated near tip stress enlarges the population of
hydrogen in the tip area [e.g. 49-55].
Finite element analysis within the framework of continuum elasto-plastic
theory shows that for an emitting crack with large plastic deformation around the crack
tip, the accumulation of hydrogen is related to the population of defects associated with
plastic deformation [49, 52]. Most of the hydrogen is trapped close to the blunted tip,
the total population of hydrogen is dominated by the density of the deep traps, which
can rise two orders of magnitude as the strain near the tip increases from zero to 0.8 and
then saturates [49]. In this case, the solubility of hydrogen is increased because of the
increase in the deep trapping sites [56]. When limited plastic strain precedes the fracture
event, most of the hydrogen resides at the normal interstitial lattice sites at the
hydrostatic stress peak located away from the tip [54, 55]. The concentration of
hydrogen is increased within the solubility limit. The peak concentration of hydrogen is
higher than the far field value only by a factor of 2~3, instead of the several orders of
magnitude predicted by elastic models. This continuum elasto-plastic model may not be
valid for a non-emitting crack. In the presence of the elastic enclave at a non-emitting
crack tip, the maximum tensile stress near the tip is not 3 to 5 times of the yield strength
located a distance away from the tip, as predicted by continuum elasto-plastic
mechanics. When Griffith cleavage occurs, the normal stress within the cohesive zone
reaches the theoretical bond strength, resulting in an enrichment of hydrogen population
of several orders of magnitude [51]. In this case, predictions of elastic interaction
models may apply and the chemical potential, , in (15) should be replaced by + ,
where is the interaction energy between hydrogen atoms and the tip stress field.
Within the framework of linear fracture mechanics the elastic interaction energy has the
form

with

under plane strain conditions in a polar coordinate system (r, ) with the origin at the
crack tip. Here KI i s the mode I stress intensity factor at the crack tip, and is the
elastic relaxation volume of the solute atom. For a cleavage crack on Fe(100) plane
within the cohesive zone, = 12.4 kJ/mol, which is about one third of the chemical
potential of hydrogen in solid solution at fugacity f =300 bar. The enhancement of the
tip stress field is significant.

2
This by no means excludes the existence of a dislocation free zone (DFZ). Deferring from the elastic core
around a non-emitting tip, DFZ around an emitting crack tip is a zone through which dislocations nucleated
from the tip can pass but otherwise is dislocation free.
44 J.-S. WANG

6. Summary

The present thermodynamic and kinetic analyses show that in a hydrogen-containing


single crystal, segregation of hydrogen from solid solution to the crack tip and the
fracture surfaces reduces the cohesive energy of the lattice, c . Hydrogen in solute
solution may also reduce the critical energy release rate for dislocation nucleation from
the crack tip, Gdisl . When the reduction in c is greater than the reduction in Gd i s l ,
hydrogen induces embrittlement. In the case where the effect of hydrogen on G disl could
be neglected the maximum propensity for embrittlement is determined by segregation
thermodynamics. The embrittlement intensity during separation is governed by the
kinetics of segregation, which is controlled by diffusion of hydrogen from the interior of
the crystal to the cohesive zone at the crack tip. The model predictions for the trends of
the temperature and cracking rate dependence of HIE in iron single crystals are
consistent with experimental observations. The model predicts that the maximum
embrittlement propensity for a single crystal is greater than that for polycrystalline
specimens. Under the same conditions the embrittlement intensity for a single crystal is
less than that for polycrystalline specimens because of the kinetics.

Acknowledgement

This work was finished while the author was supported by the Industrial Consortium of
Coating-by-Design at Northwestern University and the IHPTET Fiber Development
Consortium of DARPA. The author would like to thank Dr. Richard Hoffman for his
valuable suggestions and help in preparing the manuscript.

References

[1]. Vehoff H. and Rothe W., Gaseous hydrogen embrittlement in FeSi- and Ni-
single crystals, Acta metall., 1983, 31, 1781-1793.
[2]. Vehoff H. and Klameth H.-K., Hydrogen embrittlement and trapping at crack
tips in Ni-single crystals, Acta metall., 1985, 33, 955-962.
[3]. Miyata K; Igarashi M., Effect of ordering on hydrogen embrittlement of Ni-Cr
alloy, Mater Trans JIM 1996, 37, 703-710
[4]. Walston W.S; Bernstein I.M; Thompson A.W., The effect of internal hydrogen
on a single-crystal nickel-base superalloy, Met Trans A, 1992, 23A, 1313-
1322.
[5]. Vehoff H., and Neumann P., Crack propagation and cleavage initiation in Fe-
2.6%Si single crystals under controlled plastic crack tip opening rate in various
gaseous environments, Acta metall., 1980, 28, 265-272.
[6]. Vehoff H., Rothe W., and Neumann P., The influence of gaseous hydrogen on
the fracture process in Fe-2.6%Si single crystals, in Proceedings of ICF5, Vol.
1, pp. 265-271, Cannes, 1981.
[7]. Katz Y; Chen X.; Lii M.J.; Lanxner M.; Gerberich W.W., The anisotropic
nature of local crack stability in bcc crystals, Eng Frac Mech, 1992, 41, 541-
567.
[8]. Terasaki F; Kawakami T; Yoshikawa A; Takano N., Mechanism of crack
propagation due to hydrogen embrittlement in iron single crystals stressed
along [001] axis, Rev de Metall-Cahiers D, 1998, 95, 1519-1529.
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 45

[9]. Hu Z; Fukuyama S; Yokogawa K; Okamoto S., Hydrogen embrittlement of a


single crystal of iron on a nanometre scale at a crack tip by molecular
dynamics, Model Simul Mater Sci Eng 1999, 7, 541-551.
[10]. Magnin T; Chambreuil A; Bayle B., The corrosion-enhanced plasticity model
for stress corrosion cracking in ductile fcc alloys, Acta mater, 1996, 44, 1457-
1470.
[11]. Takasugi T; Hanada S., The influence of residual hydrogen and moisture-
released hydrogen on the embrittlement of Ni3 (Al, Ti) single-crystals, Acta
metall mater 1994, 42, 3527-3534.
[12]. Birnbaum H. K. and Sofronis P., Hydrogen-enhanced localized plasticity - a
mechanism for hydrogen-related fracture, Mater Sci & Eng, 1994, A176, 191-
202.
[13]. Sofronis P., and Birnbaum H. K., Mechanics of the hydrogen-dislocation-
impurity interactions .1. increasing shear modulus, J. Mech Phys Solids, 1995,
43, 49-90.
[14]. Ferreira P. J., Robertson I. M., and Birnbaum H. K., Hydrogen effects on
interaction between dislocations, Acta mater. 1998, 46, 1749-1757.
[15]. Anderson P. M., Wang J.-S. and Rice J. R., Thermodynamic and mechanical
models of interfacial embrittlement, in Innovation in Ultrahigh Strength Steel
Technology, ed. G. B. Olson, M. Azrin, and E. S. Wright, 1990, pp. 619-649.
[16]. McMahon C. J., Hydrogen embrittlement of high-strength steels, in Innovation
in Ultrahigh Strength Steel Technology, ed. G. B. Olson, M. Azrin, and E. S.
Wright, 1990, pp. 597-618.
[17]. Tromans D., On surface energy and the hydrogen embrittlement of iron and
steels, Acta metall mater., 1994, 42, 2043-49.
[18]. Vehoff H., Hydrogen related material problems, in Topics in Applied Physics,
Vol. 73, Hydrogen in metals III, 1997, pp. 215-278.
[19]. Rice J. R., Hydrogen and interfacial cohesion, in Effect of Hydrogen on
Behavior of Metals, ed. A. M. Thompson and I. M. Bernstein, TMS-AIME,
New York, 1976, pp.455-466.
[20]. Hirth J. P., and Rice J. R., On the thermodynamics of adsorption at interface as
it influences decohasion, Met. Trans. 11A, 1502, 1980.
[21]. Rice J. R. and Wang J.-S., Embrittlement of interfaces by solute segregation,
Mat Sci Eng, 1989, A107, 23-40.
[22]. Wang J.-S., Hydrogen induced embrittlement and the effect of the mobility of
hydrogen atoms, in Hydrogen Effects in Materials, ed. A. W. Thompson, and
N. R. Moody TMS, Warrendale, 1996, pp. 61-75.
[23]. Seah M. P., Segregation and the strength of grain boundaries, Proc. R. Soc.
Lond. 1976, A345, 535-554.
[24]. Rice J. R. and Thomson R., Ductile versus brittle behavior of crystals, Phil.
Mag., 1974, 29, 73-97.
[25]. Schoeck G., The formation of dislocation loops at crack tip in three
dimensions, Phil. Mag, 1991, A 63, 111-120.
[26]. Rice J. R., Dislocation nucleation from a crack tip - an analysis based on the
peierls concept . J. Mech Phys Solids, 1992, 40, 239-271.
[27]. McLean D., Grain Boundaries in Metals, Oxford Univ. Press, Oxford, 1957.
[28]. Hondros E. D. and Seah M. P., Grain boundary segregation, Proc. R. Soc.
Rond., 1973, A335, 191-212.
46 J.-S. WANG

[29]. Heinz K and Hammer L., Hydrogen on metals: adsorption sites and substrate
reconstruction, Phys. stat. sol. (a), 1997, 159, 225-233.
[30] Rice J. R., Beltz G. B. and Sun Y., Peierls Framework for dislocation
nucleation from a crack tip, in Topics in Fracture and Fatigue, ed. A. S. Argon,
Springer-Verlag, 1992, pp. 1-58.
[31]. Beltz G. E. and Rice J. R., Dislocation nucleation versus cleavage decohesion,
in Modeling the Deformation of Crystalline Solids: Physical Theory,
Application, and Experimental Comparisons, ed. T. C. Lowe, A. D. Rollett, P.
S. Follansbee and G. S. Daehn, TMS, 1991, pp. 457-489.
[32]. Beltz G. E. and Rice J. R., Dislocation nucleation at metal ceramic interfaces,
Acta metall mater, 40, S321-331, 1992.
[33]. Sun Y ., Y ., Rice J. R., and Truskinovsky L., Dislocation nucleation versus
cleavage in Ni3 Al and Ni, in High-Temperature Ordered Intermetallic Alloys,
ed. L. A. Johnson, D. T. Pope and J. O. Stiegler, Proc. MRS, Vol. 213, 1991,
pp.243-248.
[34]. Sun Y., Beltz G. E. and Rice J. R., Estimates from atomic models of tension
shear coupling in dislocation nucleation from a crack-tip, Mat Sci & Eng,
1993, A170, 67-85.
[35]. Sun Y., Atomistic aspects of dislocation/crack tip interaction, Ph.D.
Dissertation, Harvard University, Cambridge, MA, 1993.
[36]. Cheung K. S., Atomistic study of dislocation nucleation at a crack tip, Ph.D.
Dissertation, MIT, Cambridge, MA, 1990.
[37]. Abiko K., The evolution of iron, Phys stat sol, (a) 1997, 160, 285-196.
[38]. Yu J. and Rice J. R., Dislocation pinning effect of grain boundary segregated
solute atoms at a crack tip, in Interface structure, Properties and Design, eds.
M.H. Yoo, W.A.T. Clark and C.L. Briant, MRS Symposium Proc. Vol. 122,
MRS, 1988, pp.361-366.
[39]. Hoagland R. G., On the energetics of dislocation emission from a crack-tip in
nickel-containing hydrogen, J. Mater Res, 1994, 9, 1805-1819.
[40]. Kimura, H.,and H. Matsui, Scr metall., 1987, 21, 319.
[41]. Hirth J. P., Effects of hydrogen on the properties of iron and steels, Met.
Trans., 1980, 11A, 861-890.
[42]. Suo Z., Shih C. F., and Varias A. G., A theory for cleavage cracking in the
presence of plastic flow, Acta metall mater, 1993, 41, 1551-1557.
[43]. Beltz G. E., Rice J. R., Shih C. F., and Xia L., A self-consistent model for
cleavage in the presence of plastic flow, Acta mater, 1996, 44, 3943-3954.
[44]. Lipkin D. M., and Beltz G. E., A simple elastic cell model of cleavage fracture
in the presence of dislocation plascity, Acta mater, 1996, 44, 1287-1292.
[45]. Lipkin D. M., Clarke D. R., and Beltz G. E., A strain-gradient model of
cleavage fracture in plastically deforming materials, Acta mater, 1996, 44,
4051-4058.
[46] Fukai Y., and Sugimoto H., Hydrogen diffusion in metals-unsolved problems,
Defect and Diffusion Forum, 1992, 83, 87-110.
[47] Suzuki T., Namazue H., Koike S. and Haykawa H., Phys Rev Lett, 1983, 51,
798-800.
[48]. Beck W., Bockris J. OM., McBreen J. and Nanis L., Hydrogen permeation in
metals as a function of stress, temperature and dissolved concentration, Proc.
Roy. Soc. London, 1966, A290, 220-235.
HYDROGEN EMBRITTLEMENT IN Fe SINGLE CRYSTALS 47

[49]. Sofronis P. and McMeeking R. M., Numerical-analysis of hydrogen transport


near a blunting crack tip, J. Mech. phys. solids, 1989, 37, 317-350.
[50]. Chen, X.F., Foecke T., Lii M., Katz Y. and Gerberich W.W., The role of stress
state on H cracking in Fe-3%Si [001] single crystals, Eng Frac Mech, 1990, 35,
997-1017.
[51]. Zhang T.-Y., Shen H. and Hack J. E., The influence of cohesive force on the
equilibrium concentration of hydrogen atoms ahead of a crack tip in single
crystal iron, Scrpta metall mater, 1992, 27, 1605-1610.
[52]. Lufrano J; Sofronis P, Numerical analysis of the interaction of solute hydrogen
atoms with the stress field of a crack, Int. J. Solids Struct., 1996, 33, 1709-
1723.
[53]. Toribio J., The role of crack tip strain rate in hydrogen assisted cracking, Corr.
Sci., 1997, 39, 1687-1697.
[54]. Lufrano J; Sofronis P, Enhanced hydrogen concentrations ahead of rounded
notches and cracks-competition between plastic strain and hydrostatic stress,
Acta mater, 1998, 46, 1519-1526.
[55]. Sofronis P; Lufrano J., Interaction of local elastoplasticity with hydrogen:
embrittlement effects, Mater Sci Eng A, 1999, 260, 41-47.
[56]. Kiuchi K. and McLellan R. B., The solubility and diffusivity of hydrogen in
well-annealed and deformed iron, Acta metall, 1983, 31, 961-984.
This page intentionally left blank.
A COMPREHENSIVE MODEL FOR SOLID STATE SINTERING
AND ITS APPLICATION TO SILICON CARBIDE

H. RIEDEL AND B. BLUG


Fraunhofer-Institut fr Werkstoffmechanik
Whlerstr. 11
79108 Freiburg, Germany

Abstract: Previous models for partial aspects of solid state sintering and grain coarsen-
ing are combined to give a comprehensive model consisting of a set of equations. A
series of sinter forging tests with a SiC powder is carried out, and the model is success-
fully adjusted to the experimental results. The resulting activation energy for bulk
diffusion is substantially higher than activation energies reported in the literature.

1. Introduction

There are various possible reasons for applying powder metallurgical techniques to the
processing of materials. First, some materials such as ceramics, hard metals, refractory
metals and even certain polymers are difficult or impossible to melt and cast, so that
there is no practical alternative to the powder route. Second, steel parts are preferentially
made from powders, if the part geometry is complex, if high dimensional accuracy is
required, if large series are produced and if final machining must be avoided for
economical reasons. Accordingly, most sintered steel parts are made for the automotive
industry. Finally, powder metallurgy is applied, if a fine and homogeneous micro-
structure is needed, e.g. in critical and expensive parts for the aerospace industry.
Various techniques are used to form a powder compact (the 'green body') with the
desired shape. The most frequent shaping technique is probably die compaction, but cold
isostatic pressing is also often applied, e.g. for spark plugs or lambda probes. Whiteware
articles are shaped in large numbers by slip casting. In nearly all cases, the shaping
process, which is carried out at or near room temperature, is followed by sintering at high
temperature. In this step the fragile green body is transformed into a strong solid by the
formation of necks between the particles. In many materials, such as engineering
ceramics and hard metals, the density increases during sintering from the green density,
which is typically 55% of the theoretical bulk density in these materials, to 95 to 100%
density. Sintered steels, on the other hand, are usually sintered with less than 1%
shrinkage, since the density achievable for iron powders by die compaction is already
high enough (85 to 95%) to give good mechanical properties. Although strength and
ductility could be improved substantially by further densification during sintering, one
49
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 4970.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
50 H. RIEDEL AND B. BLUG

usually prefers to have no or little shrinkage, since dimensional accuracy of the parts is
considered to be more important than optimum strength and ductility.
Whenever the material undergoes an appreciable shrinkage during sintering,
distortions of the parts can be a serious problem. The warpage results from different
causes. Die compaction usually gives an inhomogeneous green density distribution,
which leads to differential shrinkage of different volume elements. Intended property
variations in gradient materials or in layered electronic circuits usually result in strong
warpage, unless the geometrical features and the sintering characteristics are carefully
balanced. Large and thin-walled parts undergo distortions due to gravity and to friction
on the support plate. Temperature gradients also play a role, expecially in connection
with microwave heating, since many materials absorb electromagnetic energy more
effectively at higher temperatures, which tends to enhance temperature gradients and
hence warpage.
Like a few others, e.g. [1-8], the group of the present authors has applied the finite
element method to predict the sinter distortions with the aim to minimize them by
appropriate process control [9-13].
This paper describes a comprehensive constitutive model for solid state sintering,
which summarizes various aspects published previously [14-21]. The model to be des-
cribed has already been implemented in the FE Code ABAQUS and has been applied to
the sintering of molybdenum cylinders [12]. A similar model is used by Kanters et al.
[22]. A corresponding liquid phase sintering model and its implementation was published
in [11,23].
The model in its present form is based on concepts developed, for example, by
Ashby [24,25] and Arzt [26], as far as sintering mechanisms are concerned, and on work
of Scherer [27], Abouaf et al. [1], Jagota and Dawson [28], and McMeeking and Kuhn
[29], as far as mechanical aspects are concerned. It combines results on second and third
stage sintering with models for grain growth in porous solids.
The second and third sintering stages are characterized by open and closed porosity,
respectively. In both stages, the pore surfaces are equilibrium surfaces, i.e. surfaces with
minimum energy or uniform mean curvature [15,16]. In the first stage, the surface of the
pore space is not yet in equilibrium, since it is still influenced by the initial shape of the
powder particles. Although appropriate models for neck growth in the first stage are
available [14], they are not embodied in the present model. Rather a purely pheno-
menological factor is used to describe particle rearrangement, which is another process in
the first sintering stage. Since rearrangement and nonequilibrium neck growth have
similar consequences for the constitutive response and since a detailed description of
nonequilibrium neck growth would increase the conceptual complexity of the whole
model, both processes are jointly described by one phenomenological factor.
Grain boundary diffusion is assumed to be the dominant transport mechanism, but
bulk diffusion through the grains is also taken into account as a parallel transport path.
Surface diffusion acts like a process in series to grain boundary diffusion.
A MODEL FOR SOLID STATE SINTERING 51

2. The Model Equations

2.1. THE GENERAL FORM OF THE CONSTITUTIVE EQUATION

Since the rates of stress-directed diffusion depend linearly on stress, the macroscopic
strain rate tensor must be a linear function of the stress tensor:

(1)

where the prime denotes the deviator, m is the mean (or hydrostatic) stress, p is a gas
overpressure which may develop in closed pores, ij is the Kronecker symbol, G and K
are shear and bulk viscosity, respectively, and s is the sintering stress, which arises from
the surface tension forces of the pores.
The densification rate is given by the trace of the strain rate tensor:

(2)

where is the relative density.


The dependences of G, K and s on temperature, density, grain size and possibly on
other internal variables will be specified by the detailed model to follow. It should be
mentioned that the relation between strain rate and stress may be nonlinear if the pore
space does not have an equilibrium surface, i.e. if surface diffusion plays a role. Chuang
et al. [30,31], for example, obtain 3 and dependences of the cavity growth rate
3/2

under tensile stresses, when the pore shape deviates significantly from an equilibrium
shape. Analogous solutions exist for neck growth during sintering, but for the present
purposes these nonlinear relations are not relevant, since they are valid only for stresses
several times greater than the sintering stress, while the stresses during sintering are
usually smaller than the sintering stress. In this range, the linear dependence, eq. (1), is
valid.

2.2. THE GENERAL FORM OF THE VISCOSITIES

The viscosities are written in the following form:

(3)

(4)
52 H. RIEDEL AND B. BLUG

with k = Boltzmann constant, T = absolute temperature, = atomic (or molecular) vol-


ume, R = grain radius, D b = grain boundary diffusion coefficient times grain boundary
thickness; D b exhibits the usual Arrhenius-type temperature dependence with activation
energy Q b and pre-exponential factor D b0. Further, k1 and g l are normalized bulk and
shear viscosities for open porosity, k2 and g 2 are normalized viscosities for closed
porosity, gives a smooth transition from open to closed porosity, and U is a factor to
describe the effect of grain rearrangement. Expressions for the normalized viscosities, for
and for U are given in Sections 2.3 to 2.6.
The viscosities and the sintering stress are calculated for the equilibrium pore
surfaces given in [15]. Figure 1 shows examples of equilibrium surfaces for open poro-
sity. Grain boundary diffusion in the approximately circular grain contact areas is the
dominant densification mechanism. At a certain density, neighboring contact areas touch
one another pinching off pore channels leading to isolated pores. The relative density at
which this transition from open to closed porosity occurs is given by the following
relation which approximates the numerical results of [15]:

c l = 1.05 0.115 (5)

with the dihedral angle defined by

(6)

where b and s are the specific energies of the grain boundary and the surface,
respectively.

Figure 1. Equilibrium grain surfaces for open porosity ( = dihedral angle, = porosity)
A MODEL FOR SOLID STATE SINTERING 53

2.3. THE CONTRIBUTIONS OF GRAIN BOUNDARY, BULK AND SURFACE


DIFFUSION TO THE BULK VISCOSITY

Grain boundary diffusion is considered to be the primary transport mechanism for


densification. Surface diffusion is needed to spread the material that flows out of a grain
boundary over the pore surface. In principle, the assumption of equilibrium pore shapes
implies that surface diffussion is infinitely fast. Finite surface diffusivities lead to
nonequilibrium pore surfaces.
However, the influence of finite, rather than infinite, surface diffusivities can be
treated by an approximate method which is based on the assumption of equilibrium pore
shapes [16-19]. One calculates the grain boundary and surface diffusion fluxes corre-
sponding to a sequence of equilibrium configurations. Then one equals the dissipation
rate associated with these fluxes to the negative rate of Gibbs free energy. In the resulting
densification rates grain boundary and surface diffusion act like electric resistors in
series. This approximation has been shown to give very accurate results compared to
numerical solutions for pore skrinkage in a two-dimensional configuration [19], and there
is no reason to assume that it is not applicable to 3D configurations.
Bulk diffusion is generally understood as a parallel path to the grain boundary/sur-
face diffusion path [24,25]. According to this understanding of the interaction between
grain boundary, surface and bulk diffusion, the normalized bulk viscosities are written in
the form

(7)

The subscript i denotes open (i = 1) and closed (i = 2) porosity. The subscripts b, s and v
denote grain boundary, surface and volume (or bulk) diffusion. The expressions for kib
and ki s are taken from [16-18], and the contribution of volume diffusion is treated
approximately as proposed by Ashby [24,25].

(8)

(9)

(10)

with the relative density , the porosity f = 1 , the abbreviation = 2(A3 + A 4 f) 2, the
surface diffusion coefficient, Ds = Ds0 exp(-Q s /Rg T) and the bulk diffusion coefficient
D v = D v0 exp(-Q v /R gT), where Rg is the gas constant; the functions of the dihedral angle,
A0 to A10 , are given in the Appendix. Further
54 H. RIEDEL AND B. BLUG

(11)

(12)

where
(13)

is the area fraction of grain boundaries covered by pores, and

(14)

whith = A8 2/3. The distinction between and b is made since during grain growth,
pores may detach from migrating grain boundaries (see Section 2.9). The volume fraction
of pores that remain on grain boundaries, f b , is given by

(15)

Here 0 describes the width of the range over which pore detachment occurs ( 0 = 1.3 is
chosen here), and f d is the porosity at which detachment occurs theoretically according to
the condition derived in [20,21]:

(16)

where d = A 8 d , and M b is the grain boundary mobility (see Section 2.9).


2/3

2.4. THE SHEAR VISCOSITY

For open porosity, shear viscosities were calculated in [16,29] to be:

g1 = l k l (17)
A MODEL FOR SOLID STATE SINTERING 55

An upper bound estimate for the ratio of shear to bulk viscosity is 1 = 0.6 [29] for freely
sliding grain boundaries, while a self-consistent estimate is 1 = 0.27 [16]. In the present
paper 1 is considered as an adjustable parameter,which is found to be 1.08.

For closed porosity [17]:

(18)

with

(19)

(20)

(21)

The dimensionless factor 2 should be 2 = 1 according to the self-consistent estimate


given in [17], but it is considered as an adjustable parameter in this paper.

2.5. INTERPOLATION BETWEEN OPEN AND CLOSED POROSITY

The transition parameter is assumed to vary from 0 to 1 in a density range from lo to


hi

(22)

(23)

(24)

with the relative density at pore closure c l from eq. (5), and the arbitrarily chosen
number 0.04 for the width of the transition range.
56 H. RIEDEL AND B. BLUG

2.6. PARTICLE REARRANGEMENT

The phenomenological term for grain rearrangement is written in the form

(25)

where 0 is the initial relative density, and the numbers 0.63 and 0.02 are chosen arbitra-
rily. The idea is that rearrangement can contribute to densification and deformation only
in the initial sintering stages. Above a certain density (here 63%, the relative density of a
random dense sphere packing) rearrangement can make no further contribution to
densification. If the parameter is zero, the rearrangement term has no influence. In the
following a relatively small, fixed value, = 0.2, is chosen, which means that the
rearrangement term is considered to be not very important.

2.7. THE SINTERING STRESS

Like the viscosities, the sintering stress is calculated by interpolating between the (nu-
merical) results for open [15] and closed porosity [17] using the transition parameter :

(26)

with

(27)

(28)

The functions of the dihedral angle, C 0 to C6, are given in the Appendix.

2.8. GAS PRESSURE

After pore closure entrapped gas can no longer escape from the pore space. If the gas
cannot diffuse through the solid and if ideal gas behavior is assumed, the overpressure in
the pore is
A MODEL FOR SOLID STATE SINTERING 57

(29)

where the subscript cl denotes the values of density, temperature an external pressure,
P ex, at the time of pore closure. In many cases the effect of gas pressure on the
densification rate is negligibly small. In the experiments on SiC described in Section 3,
the sintering stress is found to be of the order 5 MPa, whereas the gas overpressure is less
than 0.5 MPa at relative densities up to 98%.

2.9. GRAIN COARSENING

Sintering is usually accompanied by grain coarsening. The grain growth rate is described
by the classical Hillert law with two modifications:

(30)

The grain boundary mobility exhibits an Arrhenius-type temperature dependence


Mb = M b0 exp (-Q b /RgT).
The first modification of Hillerts law is expressed by the factor F d. It is introduced
to account for the fact that the powder usually does not have the steady-state grain size
distribution, which is implicit in the Hillert solution. The following form is chosen for Fd :

(31)

where R 0 is the initial average grain radius and can lie between - and 1. In this paper
= 0 is assumed, which corresponds to Hillerts law without a correction for the size
distribution.
The second modification, expressed by the factor Fp in eq. (29), arises from the drag
that pores exert on migrating grain boundaries. The specific form of Fp is taken from
[20,21]. For open porosity ( < c l ) :

(32)

For closed porosity ( > c l ):

(33)
58 H. RIEDEL AND B. BLUG

The Ds are functions of the dihedral angle (see Appendix). The term 1 D3f 1/2 in eq.
(32) represents the area fraction of grain boundary relative to the total grain surface area.
Since the pore drag model was not designed to be accurate for large porosities, this term
may erroneously become negative. If this happens, 1 D 3f 1/2 is set to a small positive
value.

2.10. SUMMARY OF THE MODEL

Equations (1) to (33) define the solid state sintering model. Equations (1) and (30) are the
most important ones, since they are the evolution equations for the strain rate and for the
grain size. The rest of the equations explains the quantities that appear in the evolution
equations.
The whole set of equations has been implemented in the FE Code ABAQUS as a
user supplied material routine (UMAT). An application to the sintering of Mo cylinders
is described in [12]. Also a Fortran program for the solution of eqs. (1) to (33) for
prescribed stresses was written. This is used to adjust the model parameters to the
experiments on SiC described next.

3. Experimental

3.1. MATERIAL AND SPECIMENS

An -silicon-carbide powder of Elektroschmelzwerk Kempten, Ekasic D, was used for


the tests. The powder has a grain size of 0.48 m and contains C, B and Al as sintering
aids (<1% each). The theoretical density of the fully dense material is 3.19 g/cm3.
For the subsequent sinter forging tests cylindrical specimens with 16 mm diameter
and 20 mm height were pressed in a die to a relative density, 0 = 52%, as measured after
the removal of the organic binder.

3.2. SINTER FORGING TESTS

The sinter forging test [32], which is also called load dilatometry, was applied to deter-
mine the constitutive response of the SiC powder compacts under sintering conditions.
The principle is that during sintering an axial compressive load is applied to the
cylindrical specimen and axial and radial strains are measured.
A commercial hot press was modified for the tests. The specimens were heated in a
CFC resistor furnace from room temperature to 2100C at a rate of 15 K/min. Then the
heating was switched off without a hold time leading to a cooling rate of 46 K/min. The
temperature of the specimen was measured with a differential pyrometer, while the
furnace temperature was controlled by a W-Re thermocouple placed slightly outside the
hot zone to avoid excessive ageing. Load was applied by a hydraulic cylinder and trans-
mitted by graphite punches coated with BN to minimize chemical reactions between the
specimen and the punches. The furnace atmosphere was Argon with ambient pressure.
A MODEL FOR SOLID STATE SINTERING 59

The axial strain was measured at the load rod outside the furnace using an LVDT.
Radial strains were transmitted by a graphite rod to a second LVDT outside the furnace.
The dimensions of each specimen were measured after the test, and the LVDT signals
were corrected accordingly if necessary. Further the strain signals were corrected for the
thermal expansion of the experimental setup measured with a fully dense SiC dummy
specimen.
Depending on the axial load the specimens either shrink or expand in radial
direction during the test. Friction between the specimen and the punches therefore causes
the specimens to assume either hour-glass or barrel-type shapes. However, the BN
coating apparently acts as a good lubricant, so that the deviations from cylindrical shapes
remain small. Therefore stress and strain are evaluated, as if the specimens were
cylindrical, but the changing diameter was taken into account.
A PC program was written to control all relevant parameters of the tests
(temperature, axial load, gas pressure). It also records these parameters and the strains.

3.3. MICROSCOPIC EVALUATION

Sintered and partly sintered specimens were cut, lapped, polished and etched or fractured
to study the evolution of the grain structure. A C++ program was written to evaluate the
grain size distribution function according to the Standard Intercept Method ASTM/E112.
Figure 2 shows a sequence of microstructures resulting from tests interrupted at various
temperatures.

Figure 2. Microstructure after sintering to (a) 1911 C (fracture surface),


(b) 2100C (polished and etched section).

The grain size distributions obtained from these pictures are consistent with the
steady-state distribution function predicted by Hillert [33]. Figure 3 shows an example
for a specimen sintered at 2100C. A log-normal distribution fits the data equally well.
During grain growth, the distribution function should evolve in a self-similar manner
according to the Hillert theory, and it actually does so to a high degree of accuracy.
60 H. RIEDEL AND B. BLUG

Figure 3. Observed grain size distribution function compared with Hillerts distribution.
Specimen sintered at 2100C.

4. Experimental Results and Comparison with the Model

Fifteen sinter forging tests were carried out under constant axial loads corresponding to
initial stresses between 0 and 24 MPa. The sintering model outlined in Section 2 was
adjusted to describe the evolution of axial and radial strains and of the grain size.
It is recommended to adjust the grain coarsening kinetics first, since it is nearly
independent of the densification process, whereas the densification rate depends strongly
on the grain size. Figure 4 shows the fit with the final parameter set shown in Table 1
below. Coarsening occurs between 125 and 140 min, which corresponds to temperatures
between 1875C and the peak temperature 2100C. The activation energy is found to be
Q m = 900 kJ/mol, but 800 or 1000 kJ/mol would also give a reasonable fit. The final
grain size at the end of the sintering cycle is nearly independent of the applied axial
stress, both in the experiments and in the model.
A MODEL FOR SOLID STATE SINTERING 61

Figure 4. Evolution of the grain size experiments and model.


The peak temperature, 2100C, is reached at 140 min.

Next the grain boundary and bulk diffusion coefficients are adjusted to describe the
densification rate at increasing temperature. It turned out that no reasonable fit of the
experimental densification curves could be obtained with the activation energies reported
in the literature, e.g. Ashby [25], who gives Q b = 557 kJ/mol and Q V = 696 kJ/mol. Figure
5 shows that with the diffusivities from [25] together with the observed coarsening
kinetics the material would reach only 55% relative density as compared to 96% in the
experiments. An increase of the pre-exponential factors would lead to a higher final
density, but to a completely unsatisfactory time (or temperature) dependence. On the
other hand, a very good fit is obtained, if higher activation energies are chosen. Figure 5
illustrates the improvement. The resulting activation energy for bulk diffusion is very
much higher than values from the literature (Q V = 1300 kJ/mol vs. 696 to 912 kJ/mol
[25,34-37]), but the fit of the data requires such a high value.
The effects of surface diffusion and grain boundary diffusion on the densification
and strain rates cannot easily be separated. Hence no attempt is made to determine the
surface diffusion coefficient from the fit. Rather, the surface diffusion coefficient is set to
a large value so that surface diffusion is effectively infinitely fast and therefore has no
influence on the densification rate. An equally plausible assumption would be to set
D s = D b which has been applied in other cases.
To adjust the value of the dihedral angle one observes that the measured densifica-
tion and strain rates exhibit a distinct drop at around 90% relative density, which is
attributed to the transition from open to closed porosity. This feature is reproduced by the
model, if the dihedral angle is set to = 70 according to eq. (5). The value of the
surface energy can, in principle, be derived from the stress dependence of the densifica-
tion rate. As this dependence is small, however, the fit is relatively insensitive to the
choice of the surface energies. A plausible value, S = 1 J/m , leads to a stress dependence
that is consistent with the experiments.
62 H. RIEDEL AND B. BLUG

Figure 5. Evolution of the density. Model predictions with activation energies from the
literature [25] and with the adjusted activation energies 600 and 1300 kJ/mol.

An important parameter to fit the axial and radial strains measured in the sinter
forging tests is the ratio of shear-to-bulk-viscosity, G/K, characterized by the parameters
1 and 2 . The G/K ratio is given by the ratio betwen volumetric and deviatoric strain
rate. To fit the data it was necessary to use 4 times larger G/K ratios than predicted by the
models that are based on freely sliding grain boundaries.
Figures 6a-d show a comparison of measured and calculated strains. Apparently all
sinter forging tests can be described rather accurately by the model with a single set of
parameters. Table 1 shows the values of the parameters.
A MODEL FOR SOLID STATE SINTERING 63

Figure 6a-c. Radial and axial strains for various axial stresses. Solid lines: measured,
dashed lines: predicted values. The peak temperature, 2100C, is reached at 140 min.
64 H. RIEDEL AND B. BLUG

Figure 6d. Radial and axial strains for axial stress z = -8.5 MPa. Solid lines: measured,
dashed lines: predicted values. The peak temperature, 2100C, is reached at 140 min.

TABLE 1. Parameters of the model

Meaning Symbol Value Unit


Green density 0 0.52
Initial grain radius R0 2.4 10-7 m
External gas pressure pex 1.0 10 5 Pa
Atomic volume 2.0710- 2 9 m
Grain boundary diffusion D b0 2.310 - 9 m/s
Qb 600 kJ/mol
Surface diffusion D s0 4.010 -2 * m/s
Qs 600 kJ/mol
Volume diffusion D v0 1.210 15 m/s
Qv 1300 kJ/mol
Grain boundary mobility bM b0 /4 3.810 5 m/s
Qm 900 kJ/mol
Surface energy s 1.0 J/m
Dihedral angle 70 degree
Exponent in rearrangement term 0.2
Initial grain size distribution 1
Pore detachment 0 1.3
G/K for open porosity 1 1.08
Multiplier for closed porosity 2 4
*) The surface diffusion coefficient is not fitted, but set to a high value
A MODEL FOR SOLID STATE SINTERING 65

5. Discussion

In Section 4 it was demonstrated that the proposed sintering model describes the obser-
ved densification, deformation and grain coarsening kinetics in remarkable detail. The
adjustment of the model to the measured data requires an activation energy for bulk
diffusion that is much higher than values reported for SiC in the literature. With the
diffusion data given by Ashby [25], for example, the material could never be sintered to a
high density, since grain coarsening would prevent densification beyond some 55% in
this case. Apparently the sintering aids in the present material (C, B, Al) promote a
densification mechanism with a high activation energy.
In general, the sintering aids have various effects on densification [38-40]. First
they suppress grain coarsening mechanisms that are found to operate in pure SiC at
moderate temperatures (e.g. 1200C), which effectively prevents densification [40]. This
kind of coarsening may be caused by the formation of elemental Si or of the volatile
oxide Si0. This would allow coarsening either by rapid surface diffusion or by the
evaporation/condensation mechanism. Apparently these low-temperature coarsening
mechanisms are suppressed in the present material, since coarsening starts only at
1875C, with the kinetics being consistent with a model based on grain boundary
migration.
Second the sintering aids in SiC are thought to enhance the diffusivities along grain
boundaries and through the grains. According to the present analysis they do so
especially at high temperature, since the activation energies required to fit the experi-
mental data are much higher than those obtained from diffusion experiments with pure
SiC.
The conclusion that the activation energy for bulk diffusion should be high is
probably independent of the specific sintering model proposed in this paper. Other
models such as that of McMeeking and Kuhn [29] may differ in detail, but lead to similar
trends. The activation energy for grain boundary diffusion resulting from the fit lies in
the expected range, but the value is not certain, since it is sensitive to the choice of the
rearrangement parameters, for which only assumed values could be used. Other choices
of these parameters lead to somewhat higher activation energies for grain boundary
diffusion.

6. Conclusions

Previous models on various stages of sintering and grain coarsening were combined to
form a comprehensive model of solid state sintering. This model consists of a set of
equations which is cast in a Fortran program for the analysis of sinter and sinter forging
tests. It is also implemented as a user defined material routine for the finite element code
ABAQUS.
A series of sinter forging tests (load dilatometry) was carried out with a SiC powder
containing C, B and Al as sintering aids. It is possible to adjust the model parameters
such that the model describes the densification, the deformation and the grain growth
66 H. RIEDEL AND B. BLUG

kinetics in all tests accurately. The model gives the general trend of the densification with
time or temperature, it reflects the distinct drop of the densification rate at the transition
from open to closed porosity, and it exhibits the observed stress dependence of the
deviatoric strain rate.
In this sense the model, especially in conjunction with the finite element method, is
well suited to describe technological sintering processes. Admittedly, the model contains
a number of adjustable parameters (primarily the diffusion coefficients) which are not
measured independently. Hence it cannot be concluded definitively that the good
agreement with the experiments supports all details of the model in a physical sense.
The adjustment of the model to the experiments unequivocally requires an unusu-
ally high activation energy for bulk diffusion. The difference is tentatively explained by
the effect of the sintering aids on the diffusion mechanisms.

References

1. Abouaf, M., Chenot, J.L., Raisson, G., and Baudin, P.: Finite element simulation of
hot isostatic pressing of metal powders, Int. J. Numer. Methods Engng. 25 (1988)
191-212.
2. Chenot, J.L., in: Numerical Methods in Industrial Forming Processes, Numiform 89,
Thompson, E.G., Wood, R.D., Zienkiewicz, O.C., and Samuelson, H., A.A.
Balkema, Rotterdam, 1989, p. 1.
3. Jagota, A., Mikeska, K.R., and Bordia, R.K.: Isotropic constitutive model for sinter-
ing particle packings, Am. Ceram. Soc. 73 (1990) 2266-2273.
4. Mori, K., Finite element simulation of nonuniform shrinkage in sintering of ceramic
powder compact, in: Numerical Methods in Industrial Forming Processes,
Numiform 92, Chenot, J.L., Wood, R.D., Zienkiewicz, O.C., A.A. Balkema, Rotter-
dam, 1992, pp. 69-78.
5. Cocks, A.C.F., and Du, Z.Z.: Pressureless sintering and hiping of inhomogeneous
ceramic compacts, Acta metall. mater. 41 (1993) 2113-2126.
6. Gillia, O., Bouvard, D., Doremus, P., and Imbault D.: Numerical simulation of
compaction and sintering of cemented carbide, in: European Conference on
Advances in Hard Materials Production, 1996, pp. 61-68.
7. Shinagawa, K.: Finite element analysis of microscopic material behavior in sintering
and prediction of macroscopic shape change in sintered bodies, in: Proceedings of
rd
3 Asia-Pacific Symposium AEPA, 1996, pp. 439-444.
8. Olevsky, E.A.:Theory of sintering: From discrete to continuum, Mater. Sci. Engng.
R23 (1998) 41-100.
9 . Riedel, H., and Sun, D.-Z.: Simulation of die pressing and sintering of powder
metals, hard metals and ceramics, in: Numerical Methods in Industrial Forming
Processes, Numiform 92, Chenot, J.L., Wood, R.D., and Zienkiewicz, O.C., Eds.,
A.A. Balkema, Rotterdam, 1992, pp. 883-886.
10. Sun, D.-Z., and Riedel, H.: Prediction of shape distortions of hard metal parts by
numerical simulation of pressing and sintering, Simulation of Materials Processing:
Theory, Methods and Applications, Numiform 95, Shen, S.-F. and Dawson, P.R.,
Balkema, Rotterdam, 1995, pp. 881-886.
A MODEL FOR SOLID STATE SINTERING 67

11. McHugh, P.E., and Riedel, H.: A liquid phase sintering model application to Si3N4
and WC-Co, Acta metall. mater. 45 (1997) 2995-3003.
12. Plankensteiner, A.F., Parteder, E., Riedel, H., and Sun, D.-Z.: Micromechanism
based finite element analysis of the sintering behavior of refractory metal parts using
ABAQUS, in: ABAQUS User World Congress, Chester, 1999, pp. 643-657.
13. Kraft, T., Riedel, H., Stingl, P., and Wittig, F.: Finite element simulation of die
pressing and sintering, Adv. Engng. Mater. 1(1999) 107-109.
14. Svoboda, J., and Riedel, H.: New solutions describing the formation of interparticle
necks in solid-state sintering, Acta metall. mater. 43 (1995) 1-10.
15. Svoboda, J., Riedel, H., and Zipse H.: Equilibrium pore surfaces, sintering stresses
and constitutive equations for the intermediate and late stages of sintering Part I:
Computation of equilibrium surfaces, Acta metall. mater. 42 (1994) 435-443.
16. Riedel, H., Zipse, H., and Svobada, J.: Equilibrium pore surfaces, sintering stresses
and constitutive equations for the intermediate and late stages of sintering Part II:
Diffusional densification and creep, Acta metall. mater. 42 (1994) 445-452.
17. Riedel, H., Kozk V., and Svoboda, J.: Densification and creep in the final stage of
sintering, Acta metall. mater. 42 (1994) 3093-3103.
18. Riedel, H., Svoboda, J., and Zipse, H.: Numerical simulation of die pressing and
sintering Development of constitutive equations, in: Powder Metallurgy World
Congress PM94, D. Francois, Ed., Les Editions de Physique Les Ulis, Paris, 1994,
pp. 663-671.
19. Svoboda, J., and Riedel, H.: Quasi-equilibrium sintering for coupled grain boundary
and surface diffusion, Acta metall. mater. 43 (1995) 499-506.
20. Svoboda, J., and Riedel, H.: Pore-boundary interactions and evolution equations for
the porosity and the grain size during sintering, Acta metall. mater. 40 (1992) 2829-
2840.
21. Riedel, H., and Svoboda, J.: A theoretical study of grain coarsening in porous solids,
Acta metall. mater. 41 (1993) 1929-1936.
22. Kanters, J., Eisele, U., and Rdel, J.: Scale dependent sintering trajectories, Acta
mater. to be published (2000).
23. Svoboda, J., Riedel, H., and Gaebel, R.: A model for liquid phase sintering, Acta.
metall. mater. 44 (1996) 3215-3226.
24. Ashby, M.F.: A first report on sintering diagrams, Acta metall. 22 (1974) 275-289.
25. Ashby, M.F.: HIP 6.0 Background reading, University of Cambridge (1990).
26. Arzt, E.: The influence of an increasing particle coordination on the densification of
spherical powders, Acta metall. 30 (1982) 1883-1890.
27. Scherer, G.W.: Sintering inhomogeneous glasses: Application to optical waveguides,
J. Non-Crystalline Solids 34 (1979) 239-256.
28. Jagota, A., and Dawson, P.R.: Micromechanical modeling of powder compacts
Unit problems for sintering and traction induced deformation, Acta metall. 36 (1988)
2551-2561 and 2563-2573.
29. McMeeking, R.M., and Kuhn, L.T.: A diffusional creep law for powder compacts,
Acta metall. mater. 40 (1992) 961-969.
30. Chuang, T.-J., and Rice, J.R.: The shape of intergranular creep cracks growing by
surface diffusion, Acta metall. 21 (1973) 1625-1628.
68 H. RIEDEL AND B. BLUG

31. Chuang, T.-J., Kagawa, K.I., Rice, J.R., and Sills, L.B.: Non-equilibrium models for
diffusive cavitation of grain interfaces, Acta metall. 27 (1979) 265-284.
32. H. Zipse and H. Riedel, The mechanical behavior of sintering powder compacts, in:
Ceramic Transactions, Vol. 51, Processing and Technology, H. Hausner, G.L.
Messing and S.-i. Hirano, Eds., American Ceramic Society, Westerville, OH, 1995,
pp. 489-493.
33. Hillert, M.: On the theory of normal and abnormal grain growth, Acta metall. 13
(1965) 227-237.
34. Hon, M.H., and Davis, R.F.: Self-diffusion of 14 C in polycrystalline -SiC, J. Mater.
Sci. 14 (1979) 2411-02421.
35. Hon, M.H., and Davis, R.F.: Self-diffusion of 30 Si in polycrystalline -SiC, J. Mater.
Sci. (1980) 2073-2080.
36. Hong, J.D., and Davis, R.F.: Self-diffusion of carbon-14 in high-purity and N-doped
-SiC single crystals, J. Am. Ceram. Soc. 63 (1980) 546-551.
37. Hong, J.D., and Davis, R.F.: Self-diffusion of silicon-30 in -SiC single crystals, J.
Mater. Sci. 16 (1981) 2485-2494.
38. Prochazka, S.: The role of B and C in the sintering of SiC, in: Special Ceramics, No.
6, P.Popper, Ed., British Ceramic Research Association, Manchester, 1975, pp. 171-
182.
39. van Rijswijk, W., and Shanefield, D.J.: Effects of carbon as a sintering aid in silicon
carbide, J. Am. Ceram. Soc. 73 (1990) 148-149.
40. Greskovich, C., and Rosolowski, J.H.: Sintering of covalent solids, J. Am. Ceram.
Soc. 59 (1976) 336-343.
A MODEL FOR SOLID STATE SINTERING 69

Appendix: Functions of the dihedral angle used in the sintering model

In the following equations is measured in radian.

A 0 = 0.014573+0.0063822 + 0.0009983

A 1 = -0.092348 - 0.028098 + 0.016495

A 2 = 0.16242 - 0.0062352 - 0.022826

A 3 = 0.5998 + 0.00533

A 4 = -1.271 + 0.4144

A 9 = A 0 + 0.32 A1 + 0.1024 A2

A 10 = (A l + 0.64 A2 )/A9

C 0 = - 4.069 + 6.557 + 0.0253

C1 = 26.75 - 42.58 + 5.986

C 2 = -51.01 + 82.12 - 18.56

C 3 = 3/2 (2 - 3 cos + cos )


70 H. RIEDEL AND B. BLUG

C4 = 3( - /6) - 2 3 cos sin ( - /6)

C5 = C 0 + 0.32 C1 + 0.1024 C2

C6 = (C 1 + 0.64 C2) / C 5
MAPPING THE ELASTIC-PLASTIC CONTACT AND ADHESION

SINISA DJ. MESAROVIC


Department of Materials Science and Engineering,
University of Virginia,
Charlottesville, VA 22903 U. S. A.

Abstract: This paper is a review of current understanding of the behavior of elastic-


plastic spheres in contact with or without adhesion. The emphasis is on recent
computational advances. The results are presented in the form of behavior maps,
which provide qualitative and quantitative description of the behavior, for a range of
parameters. The gaps in understanding of contact and adhesion are identified.

1. Introduction

The problems of elastic-plastic contact and adhesive pull-off between two smooth
non-conforming bodies have a wide range of applications, which span several orders of
magnitude of characteristic lengths. Indentation tests have been used from the
th
beginning of the 20 century to provide a quick measure of plastic properties of metals,
but the current standards are based on empirical correlations. Friction and wear of
solids depend critically on the details of the contacts between the asperities including
the adhesive pull-off. Powder compaction processes rely upon the mutual plastic
indentation of deformable particles. Response of an Atomic Force Microscope is
intrinsically tied to the cohesive forces between the tip and substrate.
Elastic-plastic contact alone is a complex process. Yet, such process is often only
an element of a more complex processes, such as friction and wear or powder
compaction. Therefore, there is a need for simple but comprehensive maps of contact
behavior, which would provide guidance and criteria for application of contact models.
The efforts in this direction until mid-eighties are summarized in the Johnsons (1985)
monograph. Recently, advances in analytical and computational modeling have been
made. The progress in understanding the different regimes of elastic-plastic contact
has been made possible by the advance of similarity solutions by Hill et al. (1989),
Biwa and Storkers (1995) and Storkers et al. (1997). They showed that, under a
restrictive set of assumptions, there exists a self-similar solution to the problem of
contact between rigid-plastic solids with smooth non-conforming surfaces. This
solution provides a second paradigm for the contact behavior (the first being the Hertz
solution for elastic contact). This enables us to construct maps of contact behavior in
71
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 7185.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
72 S. MESAROVIC

the parameter space which consists of the regions dominated by the two analytical
solutions and the regions where the analytical solution does not exist but a quantitative
analysis can still be performed by interpolation of extrapolation. Using the finite
element simulations Mesarovic and Fleck (1999, 2000) (hereafter referred to as I and
II) defined the regime where the similarity solution is valid. They built upon the
pioneering work of Johnson (1970) and refined and extended his master curves for
correlation of indentation tests to the regimes where contact radius becomes large
compared to the effective radius of curvature as well as to the contacts between
dissimilar elastic-plastic spheres.
Decohesion of a contact is a problem closely related to the one of crack growth.
This has been first exploited by Johnson et al. (1971) who used the singular crack tip
solution to model the decohesion process following elastic contact. Maugis (1992)
observed that a relatively weak cohesion requires more detailed modeling with a
cohesive zone model analogous to the Dugdale-Barenblatt crack tip model.
Greenwood (1997), Barthel (1998) and Greenwood and Johnson (1998) extended
Maugis' approach to more complex interface cohesion potentials. The advances in
understanding the elastic-plastic and plastic fracture mechanisms (Rice, 1968,
Hutchinson, 1979, Hutchinson, 1983) have not been exploited, in the context of contact
behavior, until recently. Mesarovic and Johnson (2000) (hereafter referred to as III)
the Maugis' cohesive zone and the results (I) and (II), to provide a model for
decohesion after elastic-plastic indentation in the range of parameters where contact
unloading/decohesion is predominantly elastic. Then, they used fracture mechanics
concepts to generate a semi-phenomenological decohesion mechanism map.
This paper represents an effort to provide a concise but comprehensive summary
of the recent results (I, II and III) and the contact behavior maps that have emerged
from these results, but it also provides some new insights and highlights the gaps in
current understanding.
Hereafter, contact between spheres is considered. This generalizes to contact
between smooth, locally axisymmetric, non-conforming surfaces, such that, the
variation in surface curvatures within a region of intensive deformation is sufficiently
small.

2. Analytic methods and scaling

2.1. INDENTATION

The configuration considered is shown in Figure 1. Spheres of radii R l and R 2


are pressed together with the normal force F, so that the contact radius is a and the total
overlap h.
In the Hertz elastic solution for the frictionless contact (e.g., Johnson, 1985, Hills
et al., 1993), only one out of four elastic constants of the two spheres,
E , E , v and v , enters the solution. The equivalent modulus E* is given by
l 2 1 2
ELASTIC-PLASTIC CONTACT AND ADHESION 73

(2.1)
The equivalent radius R o suffices to describe a given geometry:
1 / Ro 1 / R 1 +1 / R2 . (2.2)
If the following assumptions hold:
(i) The contact radius is sufficiently small compared to the radius of each sphere,
so that each sphere can be treated as a semi-infinite half space, and,
(ii) Strains and deformations are small, so that the spherical profile of the bodies
in contact can be approximated by a paraboloid of revolution,
Then, the overlap h, the contact radius a and the force F are related by
(2.3)

The solution for fully sticking elastic contact is given by Mossakovski (1963) and
Spence (1968).

Fig. 1 Geometry of contact


between two spheres with radii
R1 and R 2 . Contact radius a
corresponds to the overlap (of
undeformed spheres) h.

The second "pinning point" for the contact maps is the similarity solution for the
contact between two rigid-plastic spheres. The two spheres are composed of rigid
plastic, power law solids in accordance with J2 flow theory. In uniaxial tension, the
stress is related to the strain according to
= i l / m ; i=1,2 (2.4)
where sphere 1 has a reference strength 1 and sphere 2 has a reference strength 2 .
Both solids have the same strain-hardening exponent (1 m ) .
With the assumptions (i) and (ii), the indentation solution has the property of self-
similarity, i.e., the geometry, stress and strain fields at any stage of indentation can be
expressed in terms of an invariant solution (Storkers et al., 1997). Moreover, the
solution to the problem of contact between spherical bodies is a generalization of the
solution for the contact between a rigid sphere and a semi-infinite solid, and is obtained
74 S. MESAROVIC

from the latter by appropriate scaling. The scaling law is a generalization of the
Hertzian scaling. An equivalent radius, and an equivalent strength e describes the
combined strengths of the two spheres,
(2.5)
From (2.4) and the elastic moduli, and depending on the details of the elastic-
plastic coupling (I), one can define the yield stresses (elastic limits) of the two solids in
contact, y1 and y 2 , and the equivalent yield stress o :
(2.6)
The average pressure is related to the contact radius a by the power law relation

(2.7)

and the contact area is proportional to the indentation depth,

(2.8)

where the constants c 2 ( m ) and k (m ) depend on m, but are independent of


indentation depth, and of the diameters and strengths of the bodies in contact. They are
tabulated Biwa and Storkers (1995) for both frictionless and sticking contact. The
relations (2.7) and (2.8) imply that the indentation force depends upon the indentation
depth h according to

(2.9)

The similarity solution is expected to be accurate when the elastic deformation


becomes negligible (compared to the plastic deformation) in the region around and
under the contact.

2.2. ADHESIVE PULL-OFF

Continuum mechanics analysis of adhesive contact between perfectly elastic


spheres is well understood. The main results can be expressed in terms of two non-
dimensional parameters and (Johnson and Greenwood, 1997):

(2.10)

where E * and R o are defined in (2.1-2), w is the work of adhesion, and z o is the
equilibrium spacing between the surfaces in the Lennard-Jones potential and P is the
compressive contact force. The parameter is a measure of the elastic displacement
due to adhesive forces, normalized by their range of action.
ELASTIC-PLASTIC CONTACT AND ADHESION 75

Adhesion seems to add complexity to the already complex problem of elastic-


plastic contact. However, if one assumes that the unloading/decohesion process is
predominantly elastic, an analytic solution becomes readily available. Given the
contact pressure distribution from the indentation problem with the contact radius a0 ,
and a "fracture criterion", one can solve for the total contact pressure distribution for
any contact radius a a 0 . The mathematical framework for such analysis is the rigid
punch decomposition (Hill and Storkers, 1990). Any axisymmetric pressure
distribution p (r , a ) , over the circle of radius a, can be represented as superposition of
pressures arising from elementary rigid punches ( ) d , 0 a . Then, the
pressure p (r, a) and the distribution of rigid punches ( ) are related by Abels
integral equation and its solution:
(2.11)

Displacements are readily calculated from the rigid punch solution.

Fig. 2 (a) Geometry and pressure distribution of the adhesive contact. (b) Model for
cohesion.
76 S. MESAROVIC

The fracture criterion can be based on the work of adhesion only (for the singular
crack tip model) or on both, work of adhesion and opening displacement (for the
cohesive zone model, Figure 2). In the cohesive zone model the surfaces are assumed
to separate slightly in the annulus (a < r <c) in which the adhesive stress is constant
(Figure 2a). Let 0 be the clearance between two surfaces. The cohesive traction
is

(2.12)

where t o and o are adhesion parameters (Fig. 2b). The work of adhesion is
w =to o .
For the elastic ideally plastic solids, the indentation pressure distribution is
uniform (III) for a wide range of contact sizes (I, II). If the indentation pressure is
p (r, a o ) = p o , the non-dimensional parameters, which enter the analysis, are:

(2.13)

where is the height of the crown arising from the elastically

unloaded contact.

3. Elastic plastic indentation

From the analytic solutions in Sec. 2.1, it would appear that, the frictionless
contact problem in Fig. 1, depends on no more than two non-dimensional parameters:
e / E * (or o / E * ) and m. The indentation force (or average pressure) and the
indentation depth can then be computed as functions of the normalized contact radius
a / R o . For elastic ideally plastic solids ( m ) only one non-dimensional
parameter remains. Indeed, this is confirmed by numerical studies [(I) and (II)] for the
range of parameters where the assumptions (i) and (ii) (Sec. 2.1) hold [but see (I) for
the subtle effect of the Poissons ratii and the effects of the exact coupling between
elastic and plastic deformation]. However, when the "small strain" assumption ceases
to hold, i.e., when large sliding and rotations at the edge of the contact are present, at
least two new independent parameters are needed. The choice is not obvious, but the
simplest combinations seem to be to take the ratii of the two strengths and the two radii
as additional parameters. Let y1 be the lower of the two yield strengths in (2.6).
Tentatively, we define the parameters as:

(3.1)
(3.2)
ELASTIC-PLASTIC CONTACT AND ADHESION 77

The numerical analysis has been reported in (I) and (II) for the combinations
= 1 and 0, i.e. the spheres of equal strength and a deformable sphere against a rigid
sphere, and = 0, 1 and . The cases =0 and = should be recognized as
limiting cases of contact between a very large sphere and a very small one.

3.1. FRICTIONLESS CONTACT, ELASTIC IDEALLY PLASTIC SOLIDS

First, consider the results for the contact between a rigid and a deformable sphere
( = 0 ). The master curves of normalized average pressure and the

geometric ratio a/(2hR o) are shown in Figure 3. For small contact radii, both are
functions of a single non-dimensional parameter , where o = y1 . T h i s
part of the pressure master curve is the generalization of the Johnson's (1970, 1985)
analysis of the Brinell indentation ( = ) .
After the initial yield at and a /2hR increase with increasing contact
size. The regime of constant average pressure begins at aE * / R o 50, but the full
similarity solution (1) is not achieved until aE * / R o 800 , when a / 2hR becomes
constant (= c ).
For large contact radii, both and a /2hR depend on a / R o and (Fig. 3). In
all of the cases shown, the onset of the finite deformation regime is given by the failure
of the constant pressure condition (2.7). The finite deformation regime is characterized
by The softening is caused by excessive rotations and sliding at the edge
of the contact and is closely associated with the formation of the pile-up (Norbury and
Samuel, 1928). This finite rotation/sliding effect is modulated by the effect of the
relative size of the spheres in contact, so that the onset of the finite deformation regime
depends on .
Consider now the contact between the spheres of equal strength ( = 1). Now the
case = 0 is equivalent to the case = . The case of identical spheres
( = 1, = 1) is identical boundary value problem to the one of deformable sphere
against a rigid flat ( = 0, = 0 ) considered earlier and th e master curve, such as the
one in Fig. 3, is obtained by re-scaling the results for ( = 0, = 0) to account for the
definitions of R o (2.2) and h (Fig.1).
The case of contact between a sphere and a half-space, both with the same
properties ( = 0, = 1 ), is of particular interest. This has been discussed earlier by
Gampala et al. (1994). The numerical results (II) show that the smaller sphere takes
almost all the deformation in the finite deformation regime (Figure 4). This will be
discussed in Sec. 3.3.
78 S. MESAROVIC

Fig. 3 Master curves for contact between rigid and elastic ideally plastic spheres
( = 0 ). The average pressure (2.7) and the geometric ratio a / 2hR o (2.8) are
functions ofaE * / R o o for small a, and of a / R o and for large a. Note that vertical
scales for the two variables are different.

Fig. 4 Initial and deformed configuration for the contact of a sphere and a half-space
of identical material. (a) Elastic ideally plastic solids, (b) Elastic-hardening solids,
m=3.

The above results can be represented in the form of contact maps. One such map
for the case of the rigid ball indenting a half-space ( = 0, = ) is shown in Figure
5. The map covers all the regimes mentioned above: the elastic regime, the elastic-
ELASTIC-PLASTIC CONTACT AND ADHESION 79

plastic regime, the similarity regime and the finite deformation regime. It is
emphasized that only the materials with a very low yield strain can experience the
similarity regime. For most materials, a direct transition from the elastic-plastic to the
finite deformation regime is expected. Contact maps for other cases, i.e. other values
of and , are given in (II). They differ from the one in Fig. 5 only by the positions of
lines AC and CD.

Fig.5 Map of
frictionless indentation
of an elastic ideally
plastic solid, showing
regimes of deformation
mechanisms. Contours
of average pressure
(full lines) and the
geometric ratio
a /2hR o (dashed
lines) are included. The
map is based on the
finite element results for
E * / o = 3, 10, 30,
100, 250, 500, 1,000
and 10,000.

3.2. FRICTIONLESS CONTACT, ELASTIC - HARDENING SOLIDS

To generalize this approach to include strain hardening (finite m ), define the


reference stress
80 S. MESAROVIC

(3.3)
The similarity solution (2.7) now predicts that the reduced average pressure,
(3.4)
is constant. Indentation maps, similar to the one shown in Fig. 5, can now be
constructed for strain hardening solids. As the hardening becomes stronger ( m
smaller), the deformation pattern at the edge of the indenter transits from the pile-up to
the sink-in mode. The pile-up vanishes around m =5. At about the same value of m, the
onset of finite deformation regime becomes dependent on the variation of a / 2hR o
with a / R o . For m =3, the reduced average pressure remains constant up to a very
large values of a/ R o , while a / 2hR o begins to decrease at a /R o 0 . 3 . T h e
indentation map for m=3, will then be similar to the one shown in Fig. 5, but with the
line CD shifted to the right. The transition from the elastic-plastic to the similarity
regime (line AB) will also have shifted to the right. This is expected since for a given
contact radius, plastic deformation is more diffuse in a strain hardening materials then
in the elastic ideally plastic material, and the rigid-plastic similarity solution is
applicable if elastic deformation is negligible in the region around and under the
contact. The details are given in (I) and (II).

3.3. STRUCTURAL SOFTNESS, FRICTION, PRE-STRESS AND CONTACT


STRESS DISTRIBUTIONS

In the finite deformation regime, the smaller of the two contacting spheres of
identical materials bears most of the deformation (Fig. 4). This observation is
explained by the earlier (in terms of increasing a) decrease in the average pressure that
the smaller sphere can sustain (cf. the pressure master curves in Fig.3 for different
values of ). While both spheres deform during the elastic, elastic-plastic and
similarity regimes, the smaller sphere will enter the softening regime early and the
larger sphere will then undergo elastic-plastic unloading. The effect is somewhat less
pronounced if there is a strong strain hardening present (Fig. 4b). This structural
softness is yet to be quantified for a full range of material parameters and relative sizes
of the spheres in contact.
The effects of contact friction have been discussed in (I) but only for the case of
the rigid ball indenting a half-space ( = 0 , = ). In this case, the average pressure
remains constant up to a /R o = 0.7 (the end of simulation). This single result gives
some insight into the qualitative effects of the contact friction, but much remains to be
done. As an illustration of variability of the friction effect, consider the following
observations. Fist, due to the symmetry, there is no friction effect for the contact
between two identical spheres, whatever the friction law. Thus, the effects of friction
will vary with and . Second, Johnsons (1968) experiment on copper spheres,
indicate that only a part of the contact is sticking, while the rest is slipping, so that the
fully sticking contact (I) may not be sufficiently accurate model for low friction.
ELASTIC-PLASTIC CONTACT AND ADHESION 81

Pre-existing stresses in the solids affect contact behavior only in the elastic plastic
regime (I).
In the elastic perfectly plastic case, the pressure is almost perfectly uniform
throughout the contact area:
po = H , H 3 y , (3.5)
for a wide range of parameters, throughout both the similarity and finite deformation
regimes (III). Experimental measurements by Johnson (1968) on compressed spheres
of hard drawn copper confirm that the pressure distribution for a near perfectly
plastic material is very nearly uniform. The strain hardening material gives a pressure
distribution, which rises towards the edge of the contact (III). This finding is supported
by the experiments of Timothy et al., (1987) on mildly hardening lead spheres pressed
onto a pressure sensitive film, as well as with the computational results of Ogbonna et
al., (1995).

4. Adhesive pull-off

Assuming that the unloading/decohesion after elastic-plastic contact is


predominantly elastic, and using the rigid punch decomposition from Sec. 2.2, the
problem of unloading/decohesion can be solved, provided that the initial (i.e.,
indentation) pressure distribution is known. So far only the case of uniform pressure
distribution, resulting from the contact of elastic ideally plastic solids, has been
solved (III). As in the case of elastic-adhesive contact (Johnson et al. 1971), the main
feature of the solution is the unstable pull-off; the pulling force (as a function of
contact radius) has a maximum at a finite value of contact radius a . Two models
adopted from the fracture mechanics analyses are used: the singular model,
characterized by a critical stress intensity factor (or critical energy release rate), and the
cohesive zone model, characterized by the energy of adhesion and the range of
adhesive forces (see Fig. 2). The singular solution is a mathematical asymptote to the
cohesive zone solution, when the cohesive zone (c a ) becomes small compared to
the contact size a. For range of relevant values of and S (2.13), the critical pull-off
forces computed from the two models differ by 15% or less.
The solutions given in (III) are valid only for the limited range of values of and
S where the unloading/decohesion is predominantly elastic. Outside of this region, the
pull-off behavior is governed by different physical mechanisms. These are analyzed in
semi-phenomenological manner in (III). The results are summarized in Figure 6 where
the parameter space ( , S ) is divided into regions where decohesion is governed by
different physical processes. For S<1 any plastic flow on unloading will be small,
contained, and unlikely to have any significant effect on the elastic unloading profile.
If t o > p o ( S > 1), the yielding is widespread and the unloading/decohesion process
must be analyzed as elastic-plastic (top center in Figure 6). The decohesion/fracture
mechanism in this region cannot be determined a priori; both cleavage and void growth
82 S. MESAROVIC

and coalescence are possible. However, if exceeds the limiting value of 2, the plastic
zone will extend across the whole area of radius ao , such that separation takes place by
bulk plastic flow, even to the extent of forming a neck. The pull-off force is then given
2
by Pc = a H , i.e. the same as the force with which the spheres were pushed together.
Johnson (1976) suggested a similar approximate condition for ductile separation. We
note that low hardness and small contact size promote ductile separation. Bowden and
Tabor (1950) reproduced this behavior on the macro scale by pressing a 3.2 mm
diameter clean steel ball into a block of indium (H = 10 MPa). A ductile neck is
generally observed in molecular dynamics calculations of plastic contact followed by
separation, where the contact radius is of order 2 nm, e.g. Landman et al. (1992).
Precise measurements of such behavior in an Atomic Force Microscope, with a gold tip
on a gold surface, have been made by Agrait et al. (1996), where ao varied between 1
to 3 nm.
We note that, for predominantly elastic decohesion, the work of adhesion w varies
both, in its magnitude and in its physical meaning as one moves from the line
= S ( S + 1) toward the upper left corner in Fig. 6. Along the line = S ( S + 1) the
adhesion energy is approximately equal to the effective surface energy , while in the
domain of the singular solution the work of adhesion is taken to be the fracture
toughness G IC for the given interface, whereby most of it is dissipated through the
plastic work. To the right of the line = S ( S + 1) , the separation takes place with the
whole initial contact subjected to traction t o until the gap at the edge of the initial
contact reaches the value o , when the interface fails.

5 . Summary and suggestions for future research

The frictionless elastic-plastic contact of smooth, locally axisymmetric, non-


conforming solids has been modeled numerically for a wide range of parameters. The
contact maps have been developed which determine the regions of parameter space
where the contact deformation is governed by different deformation regimes: elastic,
elastic-plastic, similarity, and finite deformation regimes. The finite deformation
regime is of primary interest since it covers a largest part of the parameter space and
since analytic solutions are available for the elastic and the similarity regimes. Four
non-dimensional parameters, characterizing materials and geometry enter the solution:
e / E (or o / E ), m , and . The results are quantitative, in that the relevant
* *

quantities such as force, pressure, contact radius can be obtained from the contact maps
or master curves. The outstanding issue is the concept of structural softness, which,
while qualitatively understood, requires quantification, particularly for strain
hardening-materials.
Contact friction is also understood only qualitatively. It diminishes the pile-up at
the edge of the contact and it extends the constant average pressure regime. Thus, it
may be conjectured that the friction effect diminishes with increasing strain-hardening
ELASTIC-PLASTIC CONTACT AND ADHESION 83

rate. The details, such as the extent of slipping versus sticking portions of the contact
and tangential stress distributions are still unknown.

Fig. 6 Map of decohesion regimes in ( , S ) space. v = 1/3, p o / y = 3 .

The problem of adhesive pull-off following elastic-plastic indentation has been


solved only for the case of predominantly elastic unloading/decohesion and only for
the elastic ideally plastic solids. The extension to strain hardening solids is simple;
the methodology is the same as in (III). While the phenomenological behavior map has
been produced, the quantitative analysis of non-elastic unloading/decohesion has not
been done. This will likely require extensive numerical effort.
84 S. MESAROVIC

6. References

Agrait, N., Rubio, G. and Vieira, S. (1996) Plastic deformation in nanometre scale contacts. Langmuir 12,
pp. 4505-4509.
Barthel, E. (1998) On the description of the adhesive contact of spheres with arbitrary interaction potentials.
J. Colloid Interface Sci. 200, pp. 7-18.
Biwa, S. & Storkers, B. 1995 An analysis of fully plastic Brinell indentation. J. Mech. Phys. Solids 43, pp.
1303-1334.
Bowden, F. P. and Tabor, D. (1950) Friction and lubrication of solids, p.309. Clarendon.
Gampala, R., Elzey, D.M. and Wadley, H.N.G. (1994) Plastic deformation of asperities during consolidation
of plasma sprayed metal matrix composite monotape. Acta Metall. 42, pp. 3209-3221.
Greenwood, J. A (1997) Adhesion of elastic spheres. Proc. Roy. Soc. Lond. A453, pp. 1277-1297.
Greenwood, J. A. and Johnson, K. L. (1998) An alternative to the Maugis model of adhesion between elastic
spheres. J. Phys. D: Appl. Phys. 31, pp. 3279-3290.
Hill, R. and Storkers, B. (1990) A concise treatment of axisymmetric indentation in elasticity. In
Elasticity: Mathematical methods and applications, p. 199-210. Eason, G. and Ogden, R.W., eds. Ellis
Horwood, Chichester, UK.
Hill, R., Storkers, B. and Zdunek, A. B. 1989 A theoretical study of the Brinell hardness test. Proc. Royal
Soc. London A436, pp. 301-330.
Hills, D. A., Nowell, D. and Sackfield, A. (1993) Mechanics of Elastic Contact. Butterworth-Heinemann,
Oxford.
Hutchinson, J. W. (1979) A course on nonlinear fracture mechanics. Technical University of Denmark.
Department of Solid Mechanics.
Hutchinson, J. W. (1983) Fundamentals of the phenomenological theory of nonlinear fracture mechanics. J.
Appl. Mechanics 105, pp. 1042-1051 .
Johnson, K. L. (1968) An experimental determination of the contact stresses between plastically deformed
cylinders and spheres. In Engineering Plasticity, pp. 341 - 461. Heyman, J. and Leckie, F. A., eds.
Cambridge University Press.
Johnson, K.L. (1970) The correlation of indentation experiments. J. Mech. Phys. Solids 18, pp. 115-126.
Johnson, K. L. (1976) Adhesion at the contact of solids. In Theoretical and Applied Mechanics, Proc. 4th
IUTAM Congress, Ed. Koiter, p. 133. North Holland, Amsterdam.
Johnson, K. L. (1985) Contact Mechanics. Cambridge University Press.
Johnson, K. L. and Greenwood, J. A (1997) An adhesion map for the contact of elastic spheres. J. Coll.
Interface Sci. 192, pp. 326-333.
Johnson, K. L., Kendall, K. and Roberts, A. D. (1971) Surface energy and contact of elastic solids. Proc.
Roy. Soc. London A 324, pp. 301-313.
Landman, U., Luedtke, W. D. and Ringer, E. M. (1992) Molecular dynamics simulations of adhesive contact
formation and friction. In Fundamentals of Friction, Eds. Singer and Pollock. Proc. NATO Ser. E,
220, Kluwer, pp.463-510.
Maugis, D. (1992) Adhesion of spheres: the JKR-DMT transition using a Dugdale model. J. Colloid
Interface Sci. 150, pp. 243-269.
Mesarovic, S. Dj. and Fleck, N. A. (1999) Spherical indentation of elastic-plastic solids. Proc. Roy. Soc.
Lond. A 455, pp. 2707-2728.
ELASTIC-PLASTIC CONTACT AND ADHESION 85

Mesarovic, S. Dj. and Fleck, N. A. (2000) Frictionless indentation of dissimilar elastic-plastic spheres. Int.
J. Solids Structures . In press.
Mesarovic, S. Dj. and Johnson, K. L. (2000) Adhesive contact of elastic-plastic spheres. J. Mech. Phys.
Solids . In press.
Mossakovski, V.I. 1963 Compression of elastic bodies under conditions of adhesion (axisymmetric case).
PMM 27 (3), pp. 418-427.
Norbury, A. L. & Samuel, T. 1928 The recovery and sinking-in or piling-up of material in the Brinell test,
and the effects of these factors on the correlation of the Brinell with certain other hardness tests. J. Iron
Steel Institute 117(1), pp. 673-687.
Ogbonna, N., Fleck, N.A. and Cocks, A. C. F. (1995) Transient creep analysis of ball indentation. Int. J.
Mech. Sci. 37 (11), pp. 1179- 1202.
Rice J. R. (1968) Mathematical analysis in the mechanics of fracture. In Fracture Vol. II, Ed. H. Liebowitz.
Academic Press, New York.
Spence, D. A. 1968 Self similar solutions to adhesive contact problems with incremental loading. Proc.
Roy. Soc. A 305, pp. 55-80.
Storkers, B., Biwa, S., and Larsson, P.-L. (1997) Similarity analysis of inelastic contact. Int. J. Solids
Structures, 34(24), 3061-3083.
Timothy, S. P., Pearson, J. M., and Hutchings, I. M. (1987) The contact pressure distribution during plastic
compression of lead spheres. Int. J. Mech. Sci. 29 (10/11), pp. 713-719.
This page intentionally left blank.
THE CRITICAL SHEAR STRESS TO TRANSMIT A PEIERLS
SCREW DISLOCATION ACROSS A NON-SLIPPING INTERFACE

PETER M. ANDERSON
Department of Materials Science and Engineering
The Ohio State University
2041 College Road
Columbus, OH 43210-1179 U. S. A.

AND

XIAOJ.XIN
Department of Mechanical and Nuclear Engineering
Kansas State University
338 Rathbone Hall
Manhattan, KS 66506-5205 U. S. A.

Abstract: The critical resolved shear stress to transmit a screw dislocation through a
non-slipping (welded) bimaterial interface is studied as a function of the elastic mismatch
across the interface and the nonlinear shear stress-relative shear displacement relation
across the incoming and outgoing slip planes. This study extends the work of Pacheco
and Mura (1969), by using a numerical approach that incorporates a variety of slip plane
relations and by adopting a formulation by Beltz and Rice (1991) that accounts for the
finite interplanar spacing across a slip plane. The geometry is specialized to the case
when slip planes are perpendicular to the interface and numerical results are obtained for
values of mismatch, , in elastic modulus equal to 20% of the average value.
Numerical results in this regime confirm the Pacheco and Mura observation that the
critical resolved shear stress is proportional to the mismatch in elastic shear modulus. A
significant new result is that the critical resolved shear stress increases with the unstable
stacking fault energy of the slip planes, but is relatively insensitive to the maximum shear
strength of the slip planes. A simple model is constructed which adequately captures the
dependence on stacking fault energy and elastic modulus mismatch. It is with pleasure
and gratitude that this work is presented on the commemoration of the 60th birthday of
Prof. James Rice, who as Ph.D. advisor to one of the authors (PMA), instilled a sense of
enthusiasm and formalism to study dislocation-defect interactions of the type described
herein.
87
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 87105.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
88 P. M. ANDERSON AND X.J. XIN

1. Introduction

The critical resolved shear stress, *, to push a dislocation past an interface or grain
boundary is a fundamental quantity that controls macroscopic yield in many materials.
The connection between * and macroscopic yield can be understood through the
familiar Hall-Petch analysis of a pile-up of screw dislocations against an obstacle. There,
the leading dislocation in the pile-up is able to push past the obstacle if the macroscopic
shear stress, resolved onto the slip plane in the slip direction, , reaches the critical value
(Hall, 1951; Petch, 1953)

(1)

Thus, macroscopic yield can be increased by either increasing the obstacle strength * or
decreasing the spacing d between obstacles. Other parameters are the elastic shear
modulus , magnitude b of the dislocation Burgers vector, and resistance o to
dislocation motion in a material free of obstacles (d = ). Equation (1) implies that
materials with stronger obstacles have larger values of kH-P , so that the yield strength or
hardness of the material is more sensitive to a change in d.
Multilayered systems consisting of alternating A/B metallic layers show distinct
features about the strength of interfaces. First, Eqn. (1) can be fit to hardness
measurements on a given A/B multilayer system, with a constant Hall-Petch slope (kH-P)
over a layer thickness range of several microns to approximately 10nm (Shinn and
Barnett, 1992). The constant slope implies that, for a given system, interfacial strength
does not vary significantly over a large range of layer thickness. However, different
metallic systems display a large range of slopes, from 7.0GPa nm 0.5 for epitaxial Ag/Cr
multilayers to as large as 42GPa nm 0.5 for epitaxial Fe/Pt multilayers (Clemens et al.,
1999). Thus, interfacial strength appears to vary significantly among metallic systems.
At layer thickness less than 5nm, many metallic multilayered systems display a
strong departure from Eqn. (1), in that experimental hardness values increase only
modestly or even decrease with further reduction in layer thickness (Clemens et al.,
1999). This sharp change in behavior is believed to occur since the structure of
interfaces changes from a semicoherent one at larger layer thickness to a coherent one at
smaller layer thickness. Semicoherent interfaces at larger layer thickness have misfit
dislocations which appear to act as pinning sites for transmission of glide dislocations
across interfaces, and these pinning sites disappear for coherent interfaces (Rao et al.,
1995; Anderson et al., 1999a). Rao and Hazzledine (1999) estimate that misfit
dislocations in a semicoherent Al/Ni interface increase the obstacle strength by /16,
based on their Embedded Atom Method analysis of the transmission process. The same
Embedded Atom Method approach applied to coherent interfaces predicts * = 0.015
for transmission from Cu to Ni across either (111) or (100) epitaxial interfaces.
STRESS TO TRANSMIT A SCREW DISLOCATION 89

Elastic continuum approaches to dislocation transmission generally assume that


interfaces are welded, in the sense that interfaces or grain boundaries do not slip or open.
Head (1953a) concluded that a Volterra dislocation with a tight, step-function
distribution of slip is repelled from a
welded interface if the dislocation lies in
the lower modulus phase ( 1 < 2 in Fig. 1)
and it is attracted to the interface if it lies in
the larger modulus phase. Deviations from
this behavior can occur for edge
dislocations, with along the x or y-
directions in Fig. 1, if Poissons ratios 1
2 (Head, 1953b; Dundurs, 1969).
For freely-slipping interfaces, Volterra
dislocations of a screw type with along
the z-direction or edge type with along
the y-direction are always attracted to the
interface. However, an edge type with
along the x-direction is attracted to the
interface only if it lies in the larger modulus
Figure 1. The geometry of a phase. Recent models employing an
screw dislocation in material 1 at interface with continuum frictional shear
distance c from the interface. properties (Hurtado and Freund, 1998) or
Both Peierls and Volterra an interface having a linear relation
descriptions of a screw dislocation between shear traction and relative shear
are shown. displacement (Shilkrot and Srolovitz, 1998)
have been pursued as intermediate cases to
the welded and freely-slipping models. However, all of these Volterra models predict a
singular force (either attractive or repulsive) on the dislocation as it is moved toward the
interface. As such, they are not able to furnish a finite value of *.
The artificial singular force can be removed by using a Peierls description of the
dislocation, with a smoothly varying slip distribution, or by having elastic properties that
vary smoothly across a diffuse interface. Pacheco and Mura (1969) considered the first
approach and concluded that the critical stress to push a Peierls-type screw dislocation
across a welded interface from material 1 to 2 is 1*/1 0.2( 2 - 1)/( 2 + 1 ),
provided that ( 2 - 1 ) / ( 2 + 1) 1. For the case of a Cu/Ni interface, the model
predicts 1 */ 1 = 0.054, based on Cu = 54.6 GPa and Ni = 94.7 GPa (Hirth and Lothe,
1982). More recent work has extended the Pacheco and Mura model to include a
slipping interface with a sinusoidal shear traction-relative shear displacement relation
(Anderson et al., 1999b). The results indicate that slipping interfaces can generate a
significantly larger * than welded ones, since they allow for spreading of the dislocation
core into the interface.
Krznowski (1991) pursued the second approach of using a diffuse interface, with an
elastic shear modulus that varies linearly from 1 to 2 over a width w. In that case, the
obstacle strength for a Volterra screw dislocation is * = ( 2 - 1 )bln(w/2b)/4 w. Thus,
90 P. M. ANDERSON AND X.J. XIN

a Cu/Ni interface with w = 20b, for example, has 1 */ 1 = 0.0067. In general, *


decreases when interfaces become more diffuse. To summarize, continuum models
suggest that interfaces are strong barriers to slip transmission when they are chemically
sharp with large, abrupt changes in elastic moduli, when they are capable of slipping
during the transmission process, and finally, when the slip profile of the transmitting
dislocation has an abrupt, step-like shape.
This manuscript extends the work of Pacheco and Mura (1969) on the critical stress
to transmit a Peierls screw dislocation, with Burgers vector parallel to the z-direction,
across a welded interface as shown in Fig. 1. One extension is that a numerical solution
for the slip distribution of the dislocation is employed, as an alternative to a two-term
Taylor series expansion used by Pacheco and Mura, in which the first term is the slip
distribution in an infinite, homogeneous material. A second extension is to account for
the finite interplanar spacing across a slip plane in the formulation of the shear
constitutive relation for that plane, as proposed by Beltz and Rice (199 1). Finally, this
analysis extends the Frenkel (1926) sinusoidal slip plane bonding relation used by
Pacheco and Mura and the bonding relation used by Xu and Argon (1995) to a more
complex description of the atomic shear traction-relative shear displacement relation
across a slip plane. One motivation to do so is that the Frenkel sinusoidal relation is too
stiff for many commonly encountered materials (Hirth and Lothe, 1982; Sun et al., 1993;
Kaxiras and Duesbery, 1993; Xu and Argon, 1995). The other motivation is that the
shear modulus, unstable stacking energy, and maximum material shear resistance can be
varied independently on the incoming and outgoing slip planes. Thus, the effect of
bonding properties across slip planes on * can be studied in detail, for the case of
welded interfaces.

2. Model Development

2.1 THE ELASTIC SHEAR STRESS ALONG A SLIP PLANE

Consider two semi-infinite, elastic phases with shear moduli 1 and 2 that are bonded
along the y-z plane as shown in Fig. 1. The phases are assumed to have the same lattice
parameters and crystal structure so that the Burgers vector, of a screw dislocation in
each phase is identical. The shear stress on the slip plane (y = 0) generated by a Volterra
screw dislocation at x =x d is provided by Head (1953a). The screw dislocation is such
that the portion of the slip plane x > xd has a relative shear displacement = u z(y = 0 +) -
u z(y = 0 -) = b across it. The portion of the slip plane x < x d has zero relative slip. In
terms of the SF/RH sign convention employed by Hirth and Lothe (1982), the line sense
of the dislocation points along the +z-direction and the vector points along the -z-
direction. If the dislocation is in material 1, at a position xd = c, then the yz-component
of shear stress along the slip plane is
STRESS TO TRANSMIT A SCREW DISLOCATION 91

(2)

and if the dislocation is in material 2, at a position xd = -c, then

(3)

Here, the mismatch in elastic shear modulus is measured by

(4)

The model of a Peierls dislocation assumes a continuous distribution d /dx of


relative slip across a slip plane of infinitesimal thickness, so that in this case

(5)

The slip distribution is defined in Fig. 1 such that in the wake of the dislocation, (x =
) = b, and ahead of the dislocation, (x = ) = 0. The yz-component of elastic shear
stress generated by this slip distribution is obtained by replacing b = (d /dx')dx' in Eqns.
(2) and (3) and integrating over the domain x = to x = ,

(6)

Included in Eqn. (6) is the superposition of a yz component of remote shear stress ,


which serves to push the dislocation from material (1) toward material (2) in Fig. 1.
Two types of normalization are used that have specific meanings, depending on the
value of x. Both el and are normalized by , where = 1 if x > 0 and = 2 if x <
0. The normalized quantities / and el / must be continuous across the interface
since the shear stress yz must be continuous due to compatibility across a non-slipping
interface. All length quantities in italics indicate normalization by b, so that x = x/b.
92 P. M. ANDERSON AND X.J. XIN

2.2 THE SHEAR STRESS-SHEAR DISPLACEMENT RELATION FOR A SLIP


PLANE

The atomic prescription for shear stress on a slip plane is expressed most simply in terms
of the relative slip between the two atomic planes on either side of the slip plane, as
shown in Fig. 2. The yz-component of shear stress S is assumed to have a periodic
dependence on and can be expressed in general by a Fourier series,

(7)

This functional form was adopted by Xu and Argon (1995).


The coefficients n specify the detailed nature of the gradual breaking and remaking
of bonds and, as such, they can take on a variety of values. Regardless, the functional
form must satisfy the elastic shear relation for the material, so that in the limit of
vanishing , h S/ = . Other important distinguishing features of the curves are us ,
the unstable stacking fault energy (Rice, 1992), and S max , the peak shear stress that
occurs during shearing of the atomic plane. us is the maximum work that can be stored
in a unit area of slip plane as it is sheared. As such, it corresponds to the maximum
positive area under a S() curve.
A three-term version of Eqn. (7) is adopted so that , u s , and Sm a x can be varied
independently. In particular, a set of parameters,

(8)

Figure 2. Geometry of the relative shear between two adjacent atomic


planes with interplanar spacing h. S denotes the local shear stress, yz ,
acting on the slip plane and = u z ( y = + h / 2 ) - u z (y=-h/2) denotes the
relative interplanar shear.
STRESS TO TRANSMIT A SCREW DISLOCATION 93

is specified for the incoming slip plane and a separate set for the outgoing slip plane.
The first three coefficients in Eqn. (7) can be determined according to

(9)

where m, the interplanar slip at which the maximum shear stress is reached, normalized
by b, is given as the root of the condition,

(10)

Compared to earlier efforts, the three-term description in Eqn. (7) provides


additional freedom to independently vary slip plane constitutive parameters. The Frenkel
(1926) sinusoidal relation is reproduced by setting 1 = 1 and 2 = 3 = 0, so that us = 1
and S max = 1. The result is shown as Curve HH in Fig. 3. As pointed out by Xu and
Argon (1995), atomistic calculations using the Embedded Atom Method (Sun et al.,
1993) and density functional theory (Kaxiras and Duesbery, 1993) show that the Frenkel
relation often overestimates us and that, typically, us < 1 and S m a x < 1 for most
materials. The Xu and Argon (1995) two-term relation is reproduced by setting 3 = 0
and prescribing and us independently. In that case, the peak shear stress is specified in
terms of u s .
Table I shows sets of ( l , 2 , 3) corresponding to a given pair ( us , S max), over the
range (1.0, 1.0) to (0.5, 0.5). In general, each pair ( us , S max) provides two values of m
that satisfy Eqn. (10) and thus two sets of ( 1 , 2 , 3) typically are shown in each entry.
However, when us < S max , the solutions to Eqn. (10) are not real and, in that case, an
entry of (-) is listed. The range for u s is extended down to 0.5, although Xu and Argon
(1995) show in a survey of slip systems in Si, Ni, Fe, and Al that us may be as small as
0.38.
Five sets of ( 1, 2 , 3) are highlighted in boldface and will be used to study the
effect of slip plane properties on transmission of a screw dislocation across a welded
interface. Set HH corresponds to ( us , S max ) = (1.0, 1.0) and yields the Frankel
sinusoidal relation as shown in Fig. 3. The label HH denotes that the parameters ( u s ,
S m a x ) are both high in value compared to other cases in Table I. Set HH2 corresponds to
the second root for (u s , S max ) = (1.0, 1.0). As such, the HH2 curve in Fig. 3 has the
same maximum positive area and the maximum shear stress as curve HH. However, the
94 P.M.ANDERSON AND X.J. XIN

TABLE I. Values of ( 1, 2 , 3 ) corresponding to values of


u s and Smax , to be used in the S() relation shown in Eqn.(7).
STRESS TO TRANSMIT A SCREW DISLOCATION 95

curve tends to rise more slowly to the maximum shear stress and, as a result, the
maximum is positioned at m = 0.33b, compared to m = 0.25b for set HH. Thus,
comparison of for sets HH and HH2 will assess the effect of bonding features other
than ( u s , S max) on interfacial barrier strength. Set MH corresponds to ( u s , S max) =
(0.8, 1.0), so that us takes on a moderate value in Table I but S m a x remains high as in
sets HH and HH2. Set MM corresponds to a two-term series as used by Xu and Argon
(1995), with ( us , S m a x) = (0.8, 0.8233) representing moderate values of each.
Comparison of the MM and MH sets will assess the effect of S m a x on . Finally, set LL
corresponds to the lowest combination of ( u s , S max ) = (0.5, 0.5) considered in Table I.
All sets follow the premise in Fig. 2 that = b/2 is a position of crystal symmetry for
which S = 0 and for which the area under the S( ) curve reaches a positive maximum
equal to u s .

Figure 3. Candidate interplanar shear stress-relative


shear displacement relations as summarized in Table I.

2.3 SOLUTION PROCEDURE

The slip profile is determined by solving the equilibrium condition,

(11)

It is understood that, in general, the atomic relation can change abruptly at the interface,
since the set (1, 2 , 3 ) 1 corresponding to a given ( u s , S max ) 1 describes the incoming
slip plane in material 1 and a second set ( l , 2, 3) 2 corresponding to ( u s , S max )2
96 P. M. ANDERSON AND X.J. XIN

describes the outgoing slip plane in material 2. A simple assumption, ( 1, 2, 3) 2 =


(1, 2, 3) 1, is proposed. The mismatch in unstable stacking energy between the
materials can then be defined by

(12)

and a comparable expression defines the mismatch, S , in S m a x of the two phases. The
results will focus entirely on the simplest case of = 1 and = 1, so that the ratio of

An important distinction in Eqn. (11) is that the elastic shear stress is based on the
distribution of slip, (x), across an infinitesimally thin plane while the atomic shear stress
is based on the distribution of slip, (x), across an interplanar thickness h as in Fig. 2.
Beltz and Rice (1991) and Rice (1992) note this distinction between infinitesimal and
interplanar slip in the treatment of dislocation emission from crack tips. We make this
distinction also and note that the expression for elastic shear stress, Eqn. (6), can be
changed to a functional dependence on (x) using

(13)

Thus, the Beltz and Rice formalism used here is distinct from the Peierls analysis, for
which (x) = (x) is assumed. For simplicity, h = b is used as a reasonable
approximation to interplanar spacing.
Pacheco and Mura (1969) obtained an approximate analytic solution to the Peierls
formalism by approximating the slip distribution in terms of a departure from that for an
isolated dislocation that is far from an interface. In the present work, the solution to Eqn.
(11) is obtained numerically by discretizing the interplanar slip distribution into N+2
intervals so that

(14)

This slip profile corresponds to a step-wise distribution in which jumps by db = b/N+1


at positions x i, i = 1 to N + 1. The governing Eqn. (11) then becomes
STRESS TO TRANSMIT A SCREW DISLOCATION 97

(15)

Thus, Eqn. (15) is a statement of N+1 equations to determine the equilibrium positions xi
for N+1 model dislocations of Burgers vector magnitude db.
This discretization addresses the Cauchy principal value of singular 1/(xi - x j) terms
by noting that the net force of a dislocation on itself is zero. The discretization in
displacement space helps to concentrate discretization points around the core of the
dislocation, where d /dx is larger. Finally, the singular interaction of the nearest model
dislocation with the interface is addressed with a singularity exclusion method discussed
in the Appendix. There, Eqn. (A4) defines i = 0 for all model dislocations i except the
one nearest to the interface.
If the applied shear stress is specified, then Eqn. (15) can furnish N+1 equations for
the N+1 unknown values of xi. This corresponds to a stress-controlled transmission of
the dislocation through the interface. A drawback to this approach is that it is difficult to
solve numerically for the unstable portion of the transmission process, when the
dislocation continues to transmit past the point of peak resistance provided by the
interface. Rather, a more stable approach is to apply a form of displacement control, in
which N is selected to be an odd number and the center of the dislocation is specified by
prescribing the position x (N+1)/2 = c. In that case, Eqn. (15) still provides N equations
for the remaining unknown xi . The equilibrium condition R(N+1)/2 = 0 for the center
model dislocation is replaced by the statement of macroscopic equilibrium that Ri = 0.
This statement furnishes the applied stress necessary to hold the entire screw dislocation
in equilibrium,

(16)

An equivalent statement of macroscopic equilibrium in the continuum framework is


obtained by multiplying both sides of Eqn. (11) by (d /dx)dx and integrating from x = -
to + ,
98 P. M. ANDERSON AND X.J. XIN

(17)

To conclude, the procedure is to use the Peierls (1940) solution to generate an initial
trial solution,

(18)
and then compute the initial R i(0) according to Eqn. (15) with applied stress given by
Eqn. (16). The new iteration on the solution is determined by

(19)

where K ij(0) is the inverse of the N x N Jacobian matrix, Ri / xj at x j = x j(0) . This


iteration process continues until the Euclidean norm of the residual ||R|| within a
specified tolerance of 10-7.

3. Results

3.1 SENSITIVITY OF THE NUMERICAL SOLUTION TO NUMBER OF GRID


POINTS

In order to test the sensitivity of the numerical solution to N, a screw dislocation in an


infinite, homogeneous material was modeled using the Peierls formalism, but with
different N. The numerical version of the Peierls formalism is to use Eqns. (15) and (16),
but with (x) = (x) so that the simplification, (1-S'/) = 1, is made in those equations.
Table II shows %Error as the maximum value of |xi - x i(0) |/|x i(0) | for a solution with a
specified N. The %Error ~ 1/N and, based on these results, N = 101 is used for
subsequent calculations with an expected accuracy of approximately 1%.

TABLE II. %Error in Peierls solution as a function of N

3.2 COMPARISON TO PRIOR WORK

Figure 4 shows a comparison of the numerical results based on the slip plane relation HH
to earlier work by Head (1953a) and Pacheco and Mura (1969). Two types of numerical
solution are shown. The first is denoted as the Beltz and Rice formalism, which uses
Eqns. (15) and (16) and second is the Peierls formalism which sets (x) = (x). In all
cases, the mismatch in elastic shear modulus is = 0.1, so that the applied stress drives
STRESS TO TRANSMIT A SCREW DISLOCATION 99

the dislocation from a material with modulus 1 into a material with larger modulus 2 =
1.22 1. The unstable stacking fault energy and maximum shear resistance have the same
relative values, so that u s (2) = 1.22 u s (1) and S max (2) = 1.22Smax (1).
For reference, the applied stress to push a Volterra dislocation to a distance c from
the interface in Fig. 1 is (Head, 1953a)

(20)

while the corresponding stress from the Pacheco and Mura (1969) solution for a Peierls
dislocation with atomic relation HH is

(21)

All solutions are nearly identical for c > 3b, but for c < b, the Head solution diverges
noticeably due to the singular nature of the Volterra approach. The Pacheco and Mura
analytic solution (solid line) and the numerical solution for the Peierls formalism (square
symbols) essentially coincide over the entire range of positions, indicating that both
solution techniques yield the same result. The numerical solution to the Beltz and Rice
formalism has a noticeably smaller applied stress to drive the dislocation over the range
0.5b < c < 2b than the Peierls approach. However, both approaches predict barrier
strengths at c = 0 that differ by less than 5%.

Figure 4. Comparison of the applied stress


versus position of the dislocation center for
different formulations.
100 P. M. ANDERSON AND X.J. XIN

3.3 THE BARRIER STRENGTH FOR DIFFERENT SLIP PLANE RELATIONS

Figure 5 shows the applied shear stress needed to push the center of the screw dislocation
to a distance c from the interface, for the 5 slip plane constitutive relations highlighted in
Table I. In all cases, the mismatch in elastic shear modulus is = 0.1 and the Beltz and
Rice formalism is used. The general trend for all cases is that the center of the
dislocation is pushed closer to the interface as the remote stress is increased. The peak
remote stress, or barrier strength , is reached when the center of the dislocation reaches
the interface.

Figure 5. Comparison of the applied stress versus


position of the dislocation center for the 5 slip plane
constitutive relations summarized in Table 1. = 0.1

and

Among the cases studied, the effect of unstable stacking energy is most important.
In particular, Cases HH and MH differ only in that HH has stacking fault energies of
1 b/2 2 and 2 b/2 2 in materials 1 and 2, but MH has stacking fault energies of
0.8 1 b/2 2 and 0.8 2b/2 2 in materials 1 and 2, respectively. This 20% reduction in
stacking fault energies produces nearly a 25% reduction in . The rationale for this
behavior will be understood when the slip profiles for these two cases are compared in
the following section.
The effect of maximum shear resistance in the slip plane constitutive relation is
secondary to that of unstable stacking fault energy. In particular, Cases MH and MM
differ only in that MH has S max (1) = 1 /2 and S max (2) = 2 /2 , but MM has S max(1) =
0.8233 1 /2 and S max (2) = 0.8233 2 /2 . This 18% reduction in maximum shear
resistance produces a modest increase (< 5%) in . A shorter range interaction also
results, in that the rise to a peak occurs closer to the interface for Case MM.
STRESS TO TRANSMIT A SCREW DISLOCATION 101

Changes in the shape of the slip plane constitutive relation appear to produce small
changes in *, provided that stacking fault energies and maximum shear resistances are
unchanged. Evidence for this stems from comparison of Cases HH and HH2, which both
have the same stacking fault energies and maximum shear resistances. These cases
display modest differences in , even up to c = 0. In contrast, * for Case LL is nearly
half that for HH and HH2. Although both the stacking fault energies and maximum
shear resistances for LL are one half those for HH and HH2, the comparisons made
earlier suggest that it is the reduction in stacking fault energy that contributes most
significantly to the reduction in *.

3.4 THE CRITICAL SLIP PROFILES FOR DIFFERENT SLIP PLANE RELATIONS

Figure 6 shows the critical dislocation slip profiles for the five cases discussed in the
previous section. Again, the mismatch in elastic shear modulus is = 0.1. The critical
configurations occur when the maximum remote stress = * is applied and the
dislocation center is pushed to the interface. It is clear that the most diffuse, spread out
dislocation core occurs for Case LL, moderate spreading occurs for MH and MM, and
the least spreading occurs for HH and HH2. The ordering of core spreading remains the
same for c 0, when the dislocation is located away from the interface at < *. The
results reflect the well-known feature of the Peierls model that the dislocation core width
diminishes as stacking fault energy increases (Hirth and Lothe, 1982). The slip
distribution for Cases MH and MM essentially overlap, as do those for Cases HH and
HH2, so that changes in the maximum shear resistance of the slip plane or features other
than unstable stacking energy appear to have little effect on the dislocation core width.

Figure 6. Comparison of the slip profiles for the 5


slip plane constitutive relations summarized in Table
I, when the center of the dislocation is at the
interface and the barrier strength, * ,is reached.

= 0.1 and and


102 P. M. ANDERSON AND X.J. XIN

The slip distribution profiles help provide an understanding of the relative


magnitudes of * in Fig. 5. It appears that * increases as the dislocation core width
diminishes or, equivalently, as the peak slope along the slip profile increases. This trend
can be observed using a simple triangular approximation for the slope d/dx of the slip
distribution as shown in Fig. 7. The maximum slope k at the center of the dislocation and
the width 2w of the dislocation must satisfy the relation kw = b. If this profile is
positioned so that the maximum slope is at the interface (i.e., c = 0 in Fig. 7), then Eqn.
(17) yields
(22)

Thus, the barrier strength for this simple profile is directly proportional to the
maximum slope in the profile. Figure 8 shows that there is reasonable agreement
between the numerical results and the prediction from Eqn. (22), with only the first term
contributing since the 5 cases considered do not have a mismatch in normalized unstable
stacking fault energy.
The barrier strength can also be correlated with the unstable stacking fault energy as
shown in Fig. 9. A linear regression analysis of the 5 numerical cases yields

(23)

Comparison of Eqns. (22) and (23) suggests that the maximum slope is proportional to
the normalized unstable stacking fault energy. However, the numerical results are
limited, in that only cases with = 0.1 and = S = 0 have been considered. Larger
values of , for example, are expected to introduce a nonlinear dependence of * on ,
so that Eqn. (23) is not valid. The nonlinear dependence on is apparent in Eqn. (22),
since d/dx| max there depends not only on us , but also . A linear dependence on
would arise only if the slip profile

Figure 8. Barrier strength * as a function of


Figure 7. An idealized slip distribution
maximum slope in the slip distribution, for
used to develop Eqn. (22).
the 5 cases highlighted in Table I. The line is
the prediction of Eqn. (22).
STRESS TO TRANSMIT A SCREW DISLOCATION 103

remained unchanged as the dislocation approached the interface, and this is expected to
occur in the limit > 0. Based on the numerical results, the correlation in Eqn. (23)
appears to be reasonable for at least 0 0.1, and additional work is needed to
determine the behavior at larger

Figure 9. Barrier strength * as a function of


normalized unstable stacking fault energy, for
the 5 cases highlighted in Table I. The line is
Eqn. (23).

4. Conclusions

The interaction of a distributed-core screw dislocation with a welded bimaterial interface


is studied numerically, using a shear constitutive relation for the incoming and outgoing
slip planes that permits the elastic shear modulus, unstable stacking fault energy, and
maximum shear resistance to varied independently. The analysis adopts a modification
to the Peierls approach as proposed by Beltz and Rice (1991), in which the finite
thickness of the slip plane is accounted for in formulating the shear constitutive relation
for the slip plane.
The mismatch in elastic modulus across the interface generates an image force on
the dislocation which tends to push the dislocation into the elastically softer medium.
The numerical results here confirm a feature suggested by Pacheco and Mura (1969) that
for small elastic mismatch, the critical stress * to push the dislocation through the
interface into the elastically stiffer material is linearly proportional to the mismatch in
elastic modulus. The results for small elastic mismatch also show that * is linearly
proportional to the normalized unstable stacking fault energy, us / b, employed in the
constitutive relation for the slip planes, but it is weakly dependent on the maximum shear
resistance of the slip planes.
104 P. M. ANDERSON AND X.J. XIN

A simple model employing a quadratic description of the slip profile provides a


useful interpretation of the numerical results. In particular, * is shown to be linearly
proportional to product of elastic modulus mismatch and maximum slope in the slip
profile. Thus, * is expected to increase for cases of larger unstable stacking fault
energy, for which dislocations have small cores, and for cases with larger mismatch in
elastic modulus. The numerical study suggests that the predictions of the simple model
are reasonable, at least for small elastic mismatch in the range of ( 2 - 1)/(2 + 1 ) <
0.1.

4. Acknowledgements

PMA and XJX acknowledge support of the Air Force Office of Scientific Research,
Grant F49620-96-1-0238 and the support of the Ohio Supercomputer Center.

5. References
Anderson, P.M., Rao, S., Cheng, Y., and Hazzledine, P.M. (1999b) The Critical Stress for Transmission of a
Dislocation Across an Interface: Results from Peierls and Embedded Atom Models, Mater. Res. Soc. Proc.
586.
Anderson, P.M., T. Foecke, and P.M. Hazzledine (1999a) Dislocation-Based Deformation Mechanisms in
Metallic Nanolaminates, MRS Bulletin 24(2), 27-33.
Beltz, G.E., and Rice, J.R. (1991) Dislocation Nucleation Versus Cleavage Decohesion at Crack Tips, in T.C.
Lowe, A.D. Rollett, P.S. Follansbee and G.S. Daehn (eds.), Modeling the Deformation of Crystalline
Solids: Physical Theory, Application and Experimental Comparisons, TMS, Warrendale, PA, pp. 457-480.
Clemens, B.M., H. Kung and S.A. Barnett (1999) Structure and Strength of Multilayers, MRS Bulletin 24(2),
20-26.
Dundurs, J. (1969) Elastic Interaction of Dislocations with Inhomogeneities, in T. Mura (ed.) Math. Theory of
Dislocations, ASME, NY, pp. 70- 115.
Eshelby, J.D. (1949) Edge Dislocations in Anisotropic Materials, Philos. Mag. 40, 903-912.
Frenkel, J. (1926) Zur Theorie der Elastizitatsgrenze und der Festigkeit Kristallinischer Korper, Z. Phys. 37,
572-609.
Hall, E.O. (1951) The Deformation and Ageing of Mild Steel: III Discussion of Results, Proc. Roy. Soc. B64,
747-753.
Head, A.K. (1953a) Edge Dislocations in Inhomogeneous Media, Proc. Phys. Soc. (London) B66, 793-801.
Head, A.K. (1953b) The Interaction of Dislocations and Boundaries, Philos. Mag. 44, 92-94.
Hirth, J.P. and Lothe, J. (1982) Theory of Dislocations, 2nd. ed., John Wiley and Sons, New York.
Hurtado, J.A. and Freund, L.B. (1998) Force on a Dislocation Near a Weakly Bonded Interface, J. Elasticity
52(2), 167-180.
Kaxiras, E., and Duesbery, M.S. (1993) Free Energies of Generalized Stacking Faults in Si and Implications for
the Brittle-Ductile Transition, Phys. Rev. Lett. 70, 3752-3755.
Krzanowski, J.E. (1991) Effect of Composition Profile Shape on the Strength of Metallic Multilayer Structures,
Scripta Metall. Mater. 25(6), 1465-1470.
Pacheco, E.S. and Mura, T. (1969) Interaction Between and Screw Dislocation and Bimetallic Interface, J.
Mech. Phys. Sol. 17, 163-170.
Peierls, R.E. (1940) The Size of a Dislocation, Proc. Phys. Soc. (London) 52, 34-37.
Petch, N.J. (1953) Cleavage Strength of Polycrystals, J. Iron Steel Inst. 174, 25-28.
Rao, S.I., P.M. Hazzledine, and D.M. Dimiduk (1995) Atomistic Simulations of Dislocation-Interface
Interactions in Metallic Nanolayers, Mater. Res. Soc. Proc. 362, 67-72.
Rao, S.I. and Hazzledine, P.M. (1999) (accepted for publication) Atomistic Simulations of Dislocation-
Interface Interactions in the Cu-Ni System, Philos. Mag. A.
STRESS TO TRANSMIT A SCREW DISLOCATION 105

Rice, J.R. (1992) Dislocation Nucleation from Crack Tips: An Analysis Based on the Peierls Concept, J. Mech.
Phys. Sol. 40, 239-271.
Shilkrot, L.E. and Srolovitz, D.J. (1998) Elastic Analysis of Finite Stiffness Bimaterial Interfaces: Application
to Dislocation-Interface Interactions, Acta Mater. 46(9), 3063-3075.
Shinn, M., Hultmann, L., and Barnett, S.A. (1992) Growth, Structure, and Microhardness of Epitaxial
TiN/NbN, J. Mater. Res. 7, 901-911.
Sun, Y., Beltz, G.E., and Rice, J.R. (1993) Estimates from Atomic Models of Tension-Shear Coupling in
Dislocation Nucleation from a Crack Tip, Mater. Sci. Eng. A 170, 67-85.
Xu, G., and Argon, A.S. (1995) Nucleation of Dislocations from Crack Tips Under Mixed Modes of Loading:
Implications for Brittle Against Ductile Behavior of Crystals, Philos Mag. A 72, 415-451.

6. Appendix

A modified exclusion method is used in the numerical method to avoid the singular
interaction between the interface and the model dislocation of Burgers vector magnitude
db that is nearest to it. Note that a Volterra screw dislocation generates a yz-component
of shear stress along the x-z plane (Hirth and Lothe, 1982)

(A1)

For a Peierls screw dislocation, the yz component of shear stress along the x-z plane is
(Eshelby, 1949)

(A2)

where = b/2. The difference in Eqns. (A1) and (A2) is small when x > b, but Eqn. (A1)
is singular at x = 0 and Eqn. (A2) is non-singular over the range of x. In addition, when
= 0, Eqns. (A1) and (A2) become identical. Thus, in order to eliminate the singularity
associated with movement of the nearest model dislocation to the interface, the
discretized form of Eqn. (11) is modified with the substitutions,

(A3)

with

db for dislocation i nearest to the interface


i = (A4)
0 otherwise

so that Eqn. (15) is produced. Final numerical results are very insensitive to the value of
in the range (0.1, 1); a value of 0.5 is adopted for the results presented.
This page intentionally left blank.
SELF-ORGANIZING NANOPHASES ON A SOLID SURFACE

Z. SUO AND W. LU
Mechanical and Aerospace Engineering Department
and Materials Institute
Princeton University
Princeton, NJ 08544

Abstract: A two-phase epilayer on a substrate may exhibit intriguing behaviors. The


phases may select stable sizes, say on the order of 10 nm. The phases sometimes order
into a periodic pattern, such as alternating stripes or a lattice of disks. The patterns may
be stable on annealing. This paper describes an irreversible thermodynamic model that
accounts for these behaviors. The phase boundary energy drives phase coarsening. The
concentration-dependent surface stress drives phase refining. Their competition may
stabilize nanoscopic phases and periodic patterns.

1. Introduction

Rice and co-workers wrote a series of papers on polycrystalline materials subject to both
stress and heat (Chuang and Rice, 1973; Chuang et al., 1979; Needleman and Rice, 1980;
Rice and Chuang, 1981). The phenomenon concerned a cavity on a grain boundary. The
cavity could change shape and size via mass transport processes (creep, diffusion on the
cavity surface, and diffusion on the grain boundary), driven by thermodynamic forces
(stress, surface energy, and grain boundary energy). Building on those of Herring
(1951), Mullins (1957), and Hull and Rimmer (1959), Rice and co-workers developed a
general approach to this complex phenomenon.
Figure 1 outlines this approach. The basic ingredients are kinematics, energetics,
and kinetics. To model an evolving structure, one first describes its configuration with
kinematic quantities: the shape of the structure, the deformation field, the concentration
field, etc. These kinematic quantities are thermodynamic coordinates. One then equips
the structure with a free energy as a functional of the kinematic quantities. The variation
of the free energy associated with the variation of the kinematic quantities defines the
driving force. Finally one provides the kinetic relations between the rate of the kinematic
quantities and the driving forces.
These ingredients are combined into a variational statement. Depending on the
type of the kinematic quantities, the variational statement ramifies into several routes to
simulate the evolution of the structure. If the kinematics comprises fields, the variational
statement leads to partial differential equations and boundary conditions. If the
kinematics comprises discrete variables, the variational statement leads to ordinary
differential equations. If the structure is divided into elements, the variational statement
leads to a finite element model.
107
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 107122.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
108 Z. SUO AND W. LU

Figure 1. A general approach to evolving structures.

In the world of dissipative, time-dependent phenomena, this approach has a long


tradition, dating back to the treatment of damped vibration by Rayleigh (1894),
irreversible thermodynamics by Prigogine (1967) and others, spinodal decomposition by
Cahn and Hilliard (1958), and heat transfer by Biot (1970). The approach, in various
forms, has been used to study diverse material structures (e.g., Khachaturyan, 1983;
Srolovitz, 1989; McMeeking and Kuhn, 1992; Gao, 1994; Freund, 1995; Suo, 1997;
Carter et al., 1997; Bower and Craft, 1998; Cocks et al., 1999). Recent applications
include self-assembled quantum dots, electromigration voids, ferroelectric domains, and
emerging crack tips; see reviews by Freund (2000) and Suo (2000). Like other fields to
which Rice has made seminal contributions, this field has advanced solid mechanics by
developing basic theories and methods to understand phenomena of practical
significance.

2. Self-Organization by Competitive Coarsening and Refining

This paper considers a particular phenomenon: self-organizing nanophases in binary


epilayers. For a decade high-resolution imaging techniques, such as Scanning Tunneling
Microscopy (STM) and Atomic Force Microscopy (AFM), have spurred intense studies
of nanoscopic activities on solid surfaces. Kern et al. (1991) deposited a submonolayer
of oxygen on a copper (110) surface. On annealing, the oxygen atoms arranged into
stripes that alternate with bare copper stripes. The width of the stripes was on the order
of 10 nm. The self-assembled nanostructure can be a template for making functional
structures. Li et al. (1999) grew ferromagnetic iron films on the oxygen-striped copper
substrate. The stripe structure was retained up to several monolayers of the iron films.
Periodic patterns have also been observed in other material systems. Pohl et al. (1999)
deposited a monolayer of silver on a ruthenium (0001) surface, and then exposed the
silver-covered ruthenium to sulfur. The epilayer became a composite of sulfur disks in a
silver matrix. The sulfur disks were of diameter about 3.4 nm, and formed a triangular
lattice.
The observations include nanoscopic phases, periodic patterns, and stability on
annealing. These behaviors are intriguing because they are absent in bulk phase
separation. The basic behaviors of bulk phase separation are well known. Below a
critical temperature, a miscible solution becomes unstable and separates into two phases.
One phase may form particles, and the other a continuous matrix. In the beginning, the
particles are small and the total area of the phase boundaries is large. The atoms at the
SELF-ORGANIZING NANOPHASES ON A SURFACE 109

phase boundaries have excess free energy. To reduce the free energy, the total area of
the phase boundaries must reduce. Consequently, atoms leave small particles, diffuse in
the matrix, and join large particles. Over time the small particles disappear, and the
large ones become larger. The process is known as phase coarsening. Time permitting,
the phases will coarsen until only one big particle left in the matrix.
The two phases usually have different atomic lattice constants. If the phase
boundaries are coherent, the lattice constant misfit induces an elastic field. For
simplicity, assume that both phases have cubic atomic lattices, with lattice constants a 1
and a 2 , respectively. The misfit strain is M = ( a1 a 2 ) / a1 . Let E be an elastic
modulus. The average elastic energy of the two-phase mixture scales as , invariant
with the particle size. Consequently, bulk elastic misfit does not stop phase coarsening.
In the preceding paragraph, we have excluded other size scales of the system.
Imagine, for example, a thin film bonded to a substrate. The film undergoes phase
separation, but the substrate does not. The film thickness provides a length scale. When
the particle size approaches and exceeds the film thickness, the total elastic energy of the
system increases with the particle size in the lateral direction (Roytburd, 1993; Pompe et
al. 1993). Consequently, the elastic energy in the film-substrate composite causes phase
refining. The two competing actionsrefining due to elasticity and coarsening due to
phase boundariescan select an equilibrium phase size. Similar competing actions are
well known in ferroelectric films and polycrystals; see Suo (1998) for review. In the
latter, the grain size provides the needed length scale. Long range interactions other than
elasticity, such as electrostatics, can also refine phases (Chen and Khachaturyan, 1993;
Ng and Vanderbilt, 1995; Ball, 1999).
Elasticity-mediated refining may account for composition modulation sometimes
observed in multi-component semiconductor films, although we cannot be certain until a
detailed model is developed and compared with experimental observations. The existing
models (Glas, 1997; Guyer and Voorhees, 1998) do not include the film thickness effect,
so they do not have the phase refining action. It would be significant to see if the model
with both coarsening and refining actions can stabilize composition modulation.
The instability of self-assembled quantum dots provides another case study; see
Freund (2000) for review. Imagine an elemental semiconductor film on another
elemental semiconductor substrate, such as germanium film on silicon substrate. The
misfit lattice constants of the two crystals induce an elastic field, assuming that the
interface is coherent. The film may break into islands to reduce the elastic energy of the
system. The shape change is via atomic diffusion on the surface. It is sometimes
observed that the islands have a narrow size distribution, and even order into periodic
patterns. Nonetheless we note that in this case elasticity cannot stop island coarsening.
Surface diffusion allows the islands to change both lateral size and height. For example,
if the islands coarsen with a self-similar shape, the average elastic energy is invariant
with the island size. Should stable, periodic islands ever be observed, something in
addition to classical elasticity must be invoked to stop coarsening.

3. A Model of a Binary Eiplayer on a Substrate

From the above discussion, it is clear that a model of self-organizing phases should
contain the following ingredients: phase separation, phase coarsening, and phase
refining. Each ingredient may be given alternative mathematical and physical
representations. We next summarize a model proposed by Suo and Lu (2000).
110 Z. SUO AND W. LU

Imagine an epilayer of two atomic species A and B on a substrate of atomic species


S. The epilayer is one atom thick, and the substrate occupies the half space x 3 < 0 ,
bounded by the x 1 -x 2 plane. The two species A and B can be both different from that of
the substrate (such as sulfur-silver on ruthenium). Alternatively, only one species of the
epilayer is different from that of the substrate (such as oxygen on copper). The epilayer
is a substitutional alloy of A and B. Atomic diffusion is restricted within the epilayer.

3.1 KINEMATICS

Two sets of kinematic quantities describe the configuration of the epilayer-substrate


composite: one for elastic deformation, and the other for mass transport. Let u i be the
displacements in the substrate. A Latin subscript runs from 1 to 3. We assume that the
epilayer is coherent on the substrate. When the substrate deforms, the epilayer deforms
by the same amount as the substrate surface. Consequently, the displacement field of the
substrate completely specifies the deformation state of the epilayer-substrate composite.
The misfit strains in the epilayer remain constant, and therefore are not represented as
thermodynamic variables.
Let C be the fraction of atomic sites on the surface occupied by species A. Imagine
a curve on the surface. When some number of A-atoms cross the curve, to maintain a flat
epilayer, an equal number of B-atoms must cross the curve in the opposite direction.
Denote the unit vector lying in the surface normal to the curve by m. Define a vector
field I in the surface (called the mass relocation), such that I m is the number of A-
atoms across a unit length of the curve. A Greek subscript runs from 1 to 2. A repeated
index implies summation. Mass conservation requires that the variation in the
concentration relate to the variation in the mass relocation as

C = I , , (1)

where is the number of atomic sites per unit area. Similarly define a vector field J
(called the mass flux), such that J m is the number of A-atoms across a unit length of
the curve on the surface per unit time. The relation between I and J is analogous to that
between displacement and velocity. The time rate of the concentration compensates the
divergence of the flux vector, namely,

C/ t = J , . (2)

3.2 ENERGETICS

We next specify the free energy as a functional of the kinematic quantities ui and C. Let
the reference state for the free energy be atoms in three unstrained, infinite, pure crystals
of A-atoms, B-atoms and S-atoms. When atoms are taken from the reference state to
form the epilayer-substrate composite, the free energy changes, due to the entropy of
mixing, the misfits among the three kinds of atoms, and the presence of the free space.
In addition, the misfits can induce an elastic field in the substrate. Let G be the free
energy of the entire composite relative to the same number of atoms in the reference
state. For an epilayer only one atom thick, we cannot attribute the free energy to
individual kinds of misfit. Instead, we lump the epilayer and the adjacent monolayers of
the substrate into a single superficial object, and specify its free energy. The free energy
of the composite consists of two parts: the bulk and the surface, namely,
SELF-ORGANIZING NANOPHASES ON A SURFACE 111

(3)

The first integral extends over the volume of the entire system, W being the elastic
energy per unit volume. The second integral extends over the surface area, being the
surface energy per unit area. Both the volume and the surface are measured in the
unstrained substrate. As a convention, we extend the value of the substrate elastic energy
W all the way into the superficial object. Consequently, the surface energy is the
excess free energy in the superficial object in addition to the value of the substrate elastic
energy. The convention follows the one that defines the surface energy for a one-
component solid.
The elastic energy per unit volume, W, takes the usual form, being quadratic in the
displacement gradient tensor, u i , j . We assume that the substrate is isotropic, with
Youngs modulus E and Poissons ratio v. The elastic energy density function is

(4)

The stresses ij are the differential coefficients, namely, W = ij u i,j .


The surface energy per unit area, , takes an unusual from. Assume that is a
function of the concentration C, the concentration gradient C, , and the displacement
gradient in the surface, u , . Expend the function (C,C, ,u, ) to the leading order
terms in the concentration gradient C, and the displacement gradient u, , namely,
(5)

where g, f and h are all functions of the concentration C. We have assumed isotropy in
the plane of the surface; otherwise both f and h should be replaced by second rank
tensors. The leading order term in the concentration gradient is quadratic because, by
symmetry, the term linear in the concentration gradient does not affect the surface
energy. We have neglected terms quadratic in the displacement gradient tensor, which
relate to the excess in the elastic constants of the epilayer relative to the substrate. We
next explain the physical content of (5) term by term.
When the concentration field is uniform in the epilayer, the substrate is unstrained,
and the function g(C) is the only remaining term in G, the excess free energy of the
composite relative to the reference state. Consequently, g(C) is the surface energy per
unit area of the composite of the uniform epilayer on the unstrained substrate. We
assume that the epilayer is a regular solution, so that the function takes the form

(6)
Here g A and g B are the excess energy of the superficial object when the epilayer is pure
A or pure B. In the special case that A, B and S atoms are all identical, gA and g B
reduce to the surface energy of an unstrained one-component solid. Due to mass
conservation, the average concentration is constant when atoms diffuse within the
112 Z. SUO AND W. LU

epilayer. Consequently, in (6) the terms involving g A and g B do not affect diffusion.
Only the function in the bracket does. The first two terms in the bracket result from the
entropy of mixing, and the third term from the energy of mixing. The dimensionless
number measures the exchange energy relative to the thermal energy kT. The g ( C)
function is convex when < 2, and nonconvex when > 2. The function is mainly
responsible for phase separation; it favors neither coarsening nor refining.
We assume that h(C) is a positive constant, h(C)= h 0 . Any nonuniformity in the
concentration field by itself increases the free energy . Consequently, the second term
in (5) is taken to represent the phase boundary energy; the term drives phase coarsening.
The first two terms in (5) are analogous to those in the model of bulk phase separation of
Cahn and Hilliard (1958). The model represents a phase boundary by a concentration
gradient field. An alternative model would represent a phase boundary by a sharp
discontinuity. The merits of the two models have been extensively discussed in the
literature, and will not be repeated here.
Now look at the last term in (5), where u 1,1 and u 2,2 are the strains in the surface.
By definition, f is the change in the surface energy per unit strain. Consequently, f
represents the residual stress in the superficial object. More precisely, it is the resultant
force per unit length. The quantity f is known as the surface stress (Cahn 1980, Rice and
Chuang 1981). The existing literature mainly concerns the surface stress for one-
component solids (Cammarata, 1994; Cammarata and Sieradzki,1994; Freund, 1998;
Gurtin and Murdoch, 1975; Willis and Bullough, 1969; Wu, 1996). In the present
problem, when the concentration is nonuniform, the surface stress is also nonuniform,
and induces an elastic field in the substrate. As stated in Section 2, such an elastic field
will refine phases. For simplicity, we assume that the surface stress is a linear function
of the concentration, f(C) = f 0 + f 1 C . Ibach (1997) has reviewed the experimental
information on this function. Surface energy can also be a function of an order
parameter. Alerhand et al. (1988) used the idea to model surface domain patterns.

3.3 KINETICS

The composite evolves by making two kinds of changes: elastic deformation in the
substrate, and mass relocation in the epilayer. Elastic deformation does not dissipate
energy, but mass transport does. Define the driving force F as the reduction of the free
energy of the composite when an atom relocates by unit distance. Following Cahn
(1961), we specify a kinetic law by relating the atomic flux linearly to the driving force:

J = M F , (7)

where M is the mobility of atoms in the epilayer. Again we have assumed isotropy in the
surface; otherwise M should be replaced by a second rank tensor.

4. Variational Statement and Partial Differential Equations

We now mix the ingredients. Recall that the driving force is defined as the reduction of
the free energy of the composite when an atom relocates by unit distance. Translating
this definition into a mathematical description, we have
SELF-ORGANIZING NANOPHASES ON A SURFACE 113

(8)

The two vector fields, u and I, are basic kinematic variables; they vary independently,
subject to no constraint. Mass transport dissipates energy, but elastic deformation does
not. The variational statement (8) embodies these considerations.
Calculate G using the equations in Section 3, giving

(9)

where
Now compare (8) and (9). The free energy variation with the mass relocation gives
the expression for the driving force for diffusion:

(10)

Because elastic deformation does not dissipate energy, the free energy variation with the
elastic displacement vanishes, leading to

(11)

in the bulk and

(12)

on the surface. Equation (11) recovers the equilibrium equation in the elasticity theory.
Equation (12) has a straightforward interpretation. Recall that the surface stress is the
resultant force (per unit length) of the residual stress in the surface. Force balance
equates the gradient of the surface stress to the tangential traction. Equation (12) sets the
boundary conditions of the elastic field in the substrate.
Observe that the last term in (5) varies with both fields u and I, and thereby couples
the two fields. The substrate displacements enter the diffusion driving force (10), and
causes the concentration field to change over time. Once concentration field changes, the
surface stress changes and, through the boundary conditions (12), alters the
displacements in the substrate.
The elastic field in a half space due to a tangential point force acting on the surface
was solved by Cerruti (see p. 69 in Johnson, 1985). A linear superposition gives the field
due to distributed traction on the surface. Only the expression u , enters the diffusion
driving force, given by

(13)
114 Z. SUO AND W. LU

The integration extends over the entire surface.


A combination of (2), (7) and (10) leads to a diffusion equation:

(14)

Equations (6), (13) and (14) define the evolution of the concentration field. Once the
concentration field is given at t = 0, these equations update it for the subsequent time.
Equation (14) looks similar to that of Cahn (1961) for spinodal decomposition. The
main difference is how elasticity is introduced. Cahn considered misfit effect caused by
composition nonuniformity in the bulk. As discussed in Section 2, such an elasticity
effect does not refine phases. Consequently, no stable pattern is expected. Indeed,
numerical simulations have shown that the phases coarsen indefinitely, limited only by
the computational cell size or computer time; see review by Chen and Wang (1996). By
contrast, the elasticity effect in our model comes from nonuniform surface stress, which
is similar to the nonuniform residual stress in the thin film discussed in Section 2. This
elasticity effect does refine phases.

5. Scales and Parameters

A comparison of the first two terms in the parenthesis in Eqn. (14) sets a length:

(15)

This length scales the distance over which the concentration changes from the level of
one phase to that of the other. Loosely speaking, one may call b the width of the phase
boundary. The magnitude of h 0 is on the order of energy per atom at a phase boundary,
namely, h 0 ~1eV . Using magnitudes h 0 ~ 10 J, ~ 10 m
19 20 2
and kT ~ 10 20 J
(corresponding to T = 700 K), we have b = 0.3 nm.
The competition between coarsening and refining (i.e., between the last two terms in
Eqn. 14) sets another length:

(16)

This length scales the equilibrium phase size. Youngs modulus of a bulk solid is about
E ~ 10 11 N/m 2 . According to the compilation of Ibach (1997), the slope of the surface
stress is on the order f1 ~ 1N/m . These magnitudes, together with h0 ~ 10 l 9 J , give
l ~ 1 0 n m . This estimate is consistent with the experimentally observed stable phase
sizes.
From (14), disregarding a dimensionless factor, we note that the diffusivity scales as
D ~ MkT/ . To resolve events occurring over the length scale of the phase boundary
width, b, the time scale is = b / D , namely,
2
SELF-ORGANIZING NANOPHASES ON A SURFACE 115

(17)

To resolve events over the length scale of the phase size, l, the time scale is
l / D = (l / b) .
2 2

Normalize the coordinates x and by b, and the time t by . In terms of the


dimensionless coordinates and time, Eqns. (13) and (14) are combined into

(18)

The system is nonlinear and nonconvex because of the function g( C). The first two
terms in (6) disappear after the differentiation in (18). The expression for g in (18),
normalized by kT , is

(19)

The problem has two dimensionless parameters: l/b a n d . The parameter l/b
measures the ratio of the equilibrium phase size to the phase boundary width; a
representative value is l/b ~ This ratio appears in front of the refining term in
(18) as a small parameter. The parameter measures the degree of the convexity of the
function g(C), which is nonconvex when > 2. Parameters describing the initial
concentration field also enter the problem. In so far as the equilibrium pattern is
concerned, only the average concentration, Cave , is important. Recall that C ave is time-
invariant because of mass conservation.

6. Linear Perturbation Analysis

This section summarizes the results of a linear perturbation analysis (Lu and Suo 1999).
As stated before, when the concentration field is uniform, the substrate is unstrained, and
the composite is in an equilibrium state. To investigate the stability of this equilibrium
state, we superpose to this uniform concentration field a perturbation of a small
amplitude. The small perturbation can be represented by a superposition of many
sinusoidal components. Consider one such component, which is a sinusoidal field in the
x 1 direction. Let C 0 be the uniform concentration from which the system is perturbed,
q 0 be the perturbation amplitude at time t = 0, and be the perturbation wavenumber.
The wavenumber relates to the wavelength as = 2 / . According to the linear
perturbation analysis, at time t the concentration field becomes

(20)
116 Z. SUO AND W. LU

Over time, the concentration field keeps the same wavenumber, but changes the
amplitude exponentially. The characteristic number is given by

(21)

with

(22)

If > 0, the perturbation amplitude grows exponentially with the time, and a
nonuniform epilayer is obtained. If < 0, the perturbation amplitude decays
exponentially with the time, and the uniform epilayer is stable.
Figure 2 plots as a function of the wavelength. We distinguish three cases:

When > 0.5 , < 0 for all wavelengths, so that the uniform epilayer is stable
against perturbation of all wavelengths.
When < 0, the curve intersects with the horizontal axis only at one point, so that
the uniform epilayer is stable for short wavelengths, but unstable for long
wavelengths.
When 0 < < 0.5 , the curve intersects with the horizontal axis at two points, so that
the uniform epilayer is stable against perturbations of long and short wavelengths,
but unstable against perturbations of an intermediate range of wavelengths. From
(22) 0 < < 0.5 means that g (C) is convex at C 0 , but is very shallow. Acting by
itself, g (C) would stabilize the uniform epilayer. In the presence of concentration-
dependent surface stress, however, the shallow convex g (C) is insufficient to
stabilize the uniform epilayer.
For a sufficiently small , the curve reaches a peak at wavelength

(23)

This wavelength corresponds to the fastest growing perturbation mode. The linear
perturbation analysis is valid so long as the perturbation amplitude q0 exp( t) is small
compared to C 0 . The results are useful to check numerical simulation. However, the
linear perturbation analysis cannot predict the equilibrium pattern, where the
concentration nonuniformity has large magnitudes. These considerations will become
clear below.
SELF-ORGANIZING NANOPHASES ON A SURFACE 117

Figure 2 The characteristic number as a function of the wavelength.

7. Numerical Simulation

As discussed in Section 5, this is a multiscale problem. In numerical simulation, the


epilayer is divided into grids. To resolve a phase boundary, the grid size should be
smaller than b, and the time step should be smaller than . Only a finite area of the
epilayer is simulated; the infinite epilayer is represented by periodic boundary conditions.
To reduce the effect of the boundary conditions on the phase pattern, the period
simulated should be much larger than l. For the diffusion process to affect events at size
scale l, the total time should be on the order of (l/b) . We have also developed
program in the reciprocal space (Chen and Shen, 1998). In addition, without using the
diffusion equation, we can minimize the free energy to obtain equilibrium phase patterns.

7.1 A SMALL SINUSOIDAL PERTURBATION AS THE INITIAL CONDITION

The parameters in this example are l/b = , = 2.6 and C ave = 0.5. The initial
-3
concentration perturbation is sinusoidal, with amplitude 10 and wavelength 2l. We
assume that the concentration field varies with x1 but not with x2 . Consequently, the
diffusion equation is one dimensional, and the substrate is in the state of plane strain
deformation. Figure 3 shows the evolving concentration field. The evolution process
appears to have three stages. In the first stage, the perturbation amplitude increases with
time exponentially, but the wavelength remains constant, as anticipated by the linear
perturbation analysis. In the second stage, stripes of a narrower width grow. This new
wavelength is selected by the fastest growth mode predicted by the linear stability
analysis. For the present parameters, Eqn. (23) gives f /l = 0.35. In the third stage, the
118 Z. SUO AND W. LU

concentration field approaches an equilibrium pattern, with stripe width about l. Energy
minimization gives the same equilibrium pattern.

Figure 3 Evolving concentration field. The initial condition is a small perturbation from a uniform field.
SELF-ORGANIZING NANOPHASES ON A SURFACE 119

7.2 A LARGE PERTURBATION AS THE INITIAL CONDITION

In this example the initial concentration field has a large-amplitude island. All the other
parameters are kept the same as before: and C ave = 0.5. Figure 4
shows the evolving concentration field. The evolution process differs conspicuously
from that of the previous example, but ends with the same equilibrium pattern.

Figure 4 Evolving concentration field in one dimension. The initial condition is a concentration island.
120 Z. SUO AND W. LU

7.3 PRELIMINARY RESULTS OF 2D CONCENTRATION FIELDS

When the concentration field is two dimensional, the nonuniform surface stress sets a
three dimensional elastic field in the substrate. Numerical simulation is time consuming.
At this writing, we have limited experience with this general situation. Figure 5 shows a
time sequence of concentration field. The basic parameters are kept the same as before.
The initial concentration field is a random perturbation of a small amplitude from the
average concentration. The computation cell size is 6.4l. The small cell size may affect
the phase pattern. It is premature to draw any conclusion from this simulation.
Nonetheless, the simulation does produce intricate patterns often seen in experiments.

Figure 5 Evolving concentration field in two dimensions.


SELF-ORGANIZING NANOPHASES ON A SURFACE 121

8. Conclusion

This paper considers a two-phase epilayer on an elastic substrate. The ingredients for
ordering a stable, nanoscopic, periodic phase pattern are identified: (i) unstable solution
for phase separation, (ii) phase boundaries for phase coarsening, and (iii) concentration-
dependent surface stress for phase refining. We include these ingredients in a Cahn-
Hilliard type model. The concentration-dependent surface stress induces an elastic field
in the substrate, which is determined by a linear superposition of the Cerruti solution.
The elastic field enters the diffusion equation, which updates the concentration field. The
results of this model available so far are surveyed, including the governing equations,
length and time scales, linear perturbation analysis, and numerical simulation. When the
concentration field is restricted within one dimension, our numerical simulations show
that the same periodic phase pattern emerges from very different initial conditions. More
simulations need be carried out for two dimensional concentration fields. The equations
can also be used to study imperfections in a nearly ordered phase pattern, such as
dislocations and domain boundaries.

Acknowledgements

This work is supported by the Department of Energy through contract (DE-FG02-


99ER45787).

References

Alerhand, O.L., Vanderbilt, D., Meade, R.D., and Joannopoulos, J.D. (1988) Spontaneous formation of stress
domains on crystal surfaces. Phys. Rev. Lett. 61,1973-1976.
Ball, P. (1999) The Self-Made Tapestry, Oxford University Press, UK.
Biot, M.A. (1970) Variational Principles in Heat Transfer, Oxford University Press, Oxford.
Bower, A.F. and Craft, D. (1998) Analysis of failure mechanisms in the interconnect lines of microelectronic
circuits. Fatigue Fracture Engineering Materials Structure 21, 611-630.
Cahn, J.W. (1961) On spinodal decomposition. Acta Metall. 9,795-801.
Cahn, J.W. (1980) Surface stress and the chemical equilibrium of small crystalsI. the case of the isotropic
surface. Acta Metall. 28, 1333-1338.
Cahn, J.W. and Hilliard, J.E. (1958) Free energy of a nonuniform system. I. interfacial free energy. J. Chem.
Phys. 28,258-267.
Carter, W.C., Taylor, J.E., and Cahn, J.W. (1997) Variational methods for microstructural-evolution theories.
JOM, 49, No. 12, pp.30-36.
Cammarata, R.C. (1994) Surface and interface stress effects in thin films. Prog. Surf. Sci. 46, 1-38.
Cammarata, R.C. and Sieradzki K. (1994) Surface and interface stresses. Annu. Rev. Mater. Sci. 24,2 1 5 - 2 3 4 .
Chen, L.-Q. and Khachaturyan A.G. (1993) Dynamics of simultaneous ordering and phase separation and
effect of long-range coulomb interactions. Phys. Rev.Lett. 70, 1477-11480.
Chen, L.-Q. and Shen J. (1998) Applications of semi-implicit Fourier-spectral method to phase field equations.
Computer Physics Communications 108, 14-158.
Chen, L.-Q. and Wang, Y. (1996) The continuum field approach to modeling microstructural evolution. JOM,
Vol. 48, No. 12, pp.13-18.
Chuang, T.-J., Kagawa, K-I., Rice, J.R., and Sills, L.B. (1979) Non-equilibrium models for diffusive
cavitation of grain interfaces. Acta. Metall. 27, 265-284.
Chuang, T.-J. and Rice, J.R. (1973) The shape of intergranular creep cracks growing by surface diffusion.
Acta. Metall. 21, 1625-1628.
Cocks, A.C.F., Gill, S.P.A., and Pan, J. (1999) Modeling microstructure evolution in engineering materials.
Advances in Applied Mechanics, 36, 81-162.
122 Z. SUO AND W. LU

Freund, L.B. (1995) Evolution of waviness on the surface of a strained elastic solid due to stress-driven
diffusion. Int. J. Solids Structures 32, 911-923.
Freund, L.B. (1998) A surface chemical potential for elastic solids. J. Mech. Phys. Solids 46, 1835-1844.
Freund, L.B. (2000) The mechanics of electronic materials. Int. J. Solids Structures. 37, 185-l96.
Gao, H. (1994) Some general properties of stress-driven surface evolution in a heteroepitaxial thin film
structure. J. Mech. Phys. Solids 42, 741-772.
Glas, F. (1997) Thermodynamics of a stressed alloy with a free surface: coupling between the morphological
and compositional instabilities. Phys. Rev. B 55, 11277-11286.
Gurtin, M.E. and Murdoch, A.I. (1975) A continuum theory of elastic material surface. Arch. Rat. Mech.
Anal. 57, 291-323.
Guyer, J.E. and Voorhees, P.W. (1998) Morphological stability and compositional uniformity of alloy thin
films. J. Crystal Growth 187, 150-165.
Herring, C. (1951) Surface tension as a motivation for sintering. The Physics of Powder Metallurgy,
McGraw-Hill, editor Kingston, W.E., New York pp. 143-179.
Hull, D. and Rimmer, D.E. (1959) The growth of grain-boundary voids under stress. Phil. Mag., 4, 673-687.
Ibach, H. (1997) The role of surface stress in reconstruction, epitaxial growth and stabilization of mesoscopic
structures. Surf. Sci. Rep. 29, 193-263.
Johnson, K.L. (1985) Contact Mechanics, Cambridge University Press, UK.
Kern, K., Niebus, H., Schatz, A., Zeppenfeld, P., George, J., Comsa, G. (1991) Long-range spatial self-
organization in the adsorbate-induced restructuring of surfaces: Cu{110}-(2x1) O. Phys. Rev. Lett. 67,
855-858.
Khachaturyan, A.G. (1983) Theory of Structural Transformation in Solids, Wiley, New York.
Li, D., Diercks, V., Pearson, J., Jiang, J.S., and Bader, S.D. (1999) Structural and magnetic studies of fcc Fe
films with self-organized lateral modulation on striped Cu{110}-O(2x1) substrates. J. Appl. Phys. 85,
5285-5287.
Lu, W. and Suo, Z. (1999) Coarsening, refining, and pattern emergence in binary epilayers, the Fred Lange
Festschrift in the journal Zeitschrift fur Metallkunde. In Press.
McMeeking, R.M. and Kuhn, L.T. (1992) A diffusional creep law for powder compacts. Acta Metall. Mater.
40, 961-969.
Mullins, W.W. (1957) Theory of thermal grooving, J. Appl. Phys., 28, 333-339.
Needleman, A and Rice, J.R. (1980) Plastic creep flow effects in the diffusive cavitation of grain boundaries.
Acta Metall. 28, 1315-1332.
Ng, K.-O. and Vanderbilt, D. (1995) Stability of periodic domain structures in a two dimensional dipolar
model. Phys. Rev. B 52, 2177-2183.
Pohl, K., Bartelt, M.C., de la Figuera, J., Bartelt, N.C., Hrbek, J., Hwang, R.Q. (1999) Identifying the forces
responsible for self-organization of nanostructures at crystal surfaces. Nature 397, 238-241.
Pompe, W., Gong, X., Suo, Z. and Speck, J.S. (1993) Elastic energy release due to domain formation in the
strained epitaxy of ferroelectric and ferroelastic films. J. Appl. Phys. 74, 6012-6019.
Prigogine, I. (1967) Introduction of Thermodynamics of Irreversible Processes, 3rd edition, Wiley, New York.
Rayleigh, J.W.S. (1894) The Theory of Sound, Vol. 1, Art. 81. Reprinted by Dover, New York.
Rice, J.R. and Chuang, T.-J. (1981) Energy variations in diffusive cavity growth. J. Am. Ceram. Soc. 64, 46-
53.
Roytburd, A.L. (1993) Elastic domains and polydomain phases in solids. Phase Transitions, 45, 1-33.
Srolovitz, D.J. (1989) On the stability of surfaces of stressed solids. Acta Metall. 37, 621-625.
Suo, Z. (1997) Motions of microscopic surfaces in materials. Advances in Applied Mechanics. 33, 193-294.
Suo, Z. (1998) Stress and strain in ferroelectrics. Current Opinion in Solid State & Materials Sicence, 3, 486-
489.
Suo, Z. (2000) Evolving materials structures of small feature sizes. Int. J. Solids Structures. 37, 367-378.
Suo, Z. and Lu, W. (2000) Composition modulation and nanophase separation in a binary epilayer, J. Mech.
Phys. Solids. In press.
Willis, J.R. and Bullough, R. (1969) The interaction of finite gas bubbles in a solid. J. Nuclear Mater. 32, 76-
87.
Wu, C.H. (1996) The chemical potential for stress-driven surface diffusion. J. Mech. Phys. Solids 44, 2059-
2077.
ELASTIC SPACE CONTAINING A RIGID ELLIPSOIDAL INCLUSION
SUBJECTED TO TRANSLATION AND ROTATION

M. KACHANOV, E. KARAPETIAN AND I. SEVOSTIANOV


Department of Mechanical Engineering,
Tufts University
Medford, MA 02155

1. Introduction

The problem of a linear elastic space containing a rigid ellipsoidal inclusion subjected to
translation and rotation is critically overviewed and further developed. Full elastic fields,
as well as the stiffness relations that give forces and moments, that have to be applied
to the inclusion in order to produce the given rotations and displacements, are given.
This problem can be viewed as supplementary to Eshelby's problem for an ellipsoidal
inhomogeneity (Eshelby, 1961) which does not cover the rigid body motion of the
inhomogeneity. The problem is of a practical interest, for example, for the geotechnical
applications (Selvadurai, 1976). In contrast with Eshelbys problem, it has not been fully
analyzed. The problem was first considered, probably, by Keer (1965) in the special case
of a rigid circular disc subjected to translation in the disc plane. The case of a general
ellipsoid was considered by Lure (1970). However, his solution is incomplete: it is not
expressed in terms of any standard functions and results for the important case of a
spheroid (that do not follow from the general case in a straightforward way) were not
given; besides, his work contains misprints and minor errors. Kanwal and Sharma (1976)
considered the case of a spheroid and derived tractions on the spheroids boundary and
relations between the overall forces and moments applied to the spheroid and its
displacements and rotations (stiffness relations). However, the full fields in the elastic
space were not given. Selvadurai (1976) considered the spheroidal inclusion subjected to
translation parallel to the spheroids axis and gave full fields in this special axisymmetric
case.
The analysis was extended to the case of the transversely isotropic space with a rigid
spheroidal inclusion in several works. Selvadurai (1979, 1980) considered this problem in
the case of a circular disk subjected to translation and to rotation about the axis of the
symmetry. Zureick (1988, 1989) considered the case of a spheroid, but the solution is not
given in the closed form. Rahman (2000) considered a rigid disc of the elliptical shape
subjected to translation normal to the disc plane and gave the closed form solution.
The present work focuses on the case of the isotropic space and further advances the
existing results, by explicitly deriving the full set of elastic fields and stiffness relations
123
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 123143.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
124 M. KACHANOV, ET AL.

in the closed form for all cases, including, in particular, the important case of a spheroid
(requiring a non-trivial limiting transition). The derivation generally follows the approach
of Lure (1970).

2. General Ellipsoid

We consider a rigid ellipsoid, with the surface given by equation

(2.1)

embedded into an infinite elastic medium. It is subjected to small translation u0 and


small rotation , both of arbitrary directions. Equation (2.1) can be rewritten in terms of
parameters (instead of a 1 , a 2 , a3)
as follows:

(2.2)

where a 0 , are ellipsoid's semiaxes. The points of the boundary


( = 0 ) undergo displacements given by

u = 0 = u0 + R 0 (2.3)

where R 0 is the position vector of a point on the ellipsoid's surface.


We use orthogonal ellipsoidal coordinates (, ,v) that are related to cartesian
coordinates ( x1 , x 2 , x 3 ) as follows:

(2.4)

where
ELASTIC SPACE WITH AN INCLUSION 125

2.1. FIELDS PRODUCED BY TRANSLATION OF AN ELLIPSOID

We utilize Papkovich-Neubers general representation of displacements in terms of


harmonic vector B and harmonic scalar B 0 :

(2.5)

where we further express B 0 and components B m of B in terms of four potential


functions F0 , F1 , F 2 , F3 as follows:

B m = C m F0 , B 0 =D m Fm , m = 1,2,3 (2.6)

and constants C m , Dm are to be determined from boundary conditions u = u0.


=0
We seek such potential functions that they are harmonic everywhere except for the
ellipsoids surface and tend to zero at . Therefore, they are taken as potentials of a
simple layer and are as follows.
Inside the ellipsoid:

F0 =1, Fm =x m , m = 1,2,3, 0 (2.7)

Outside of the ellipsoid:

(2.8)

where i ( ) are given by the elliptic integrals:

(2.9)

Utilizing (2.6-8), representation (2.5) can be written in terms of i () :


126 M. KACHANOV, ET AL.

(2.10)

Differentiating functions i ( ) :

(2.11)

brings (2.10) to the form:

(2.12)

Boundary conditions u = 0 =u0 lead to the following system of six linear algebraic
equations for constants C m , D m (that decouple into three separate systems of two
equations each):

(2.13)

Utilizing the relationship for curvilinear coordinates , where H s a r e

Lames coefficients, yields:


ELASTIC SPACE WITH AN INCLUSION 127

(2.14)

Substituting (2.14), along with expressions for constants C m ,D m obtained by solving


(2.13), into (2.12) yields the displacement field outside of the ellipsoid ( > 0 ):

(2.15)

where the following notations are used:

(2.16)

(2.17)

and where i ( ) entering (2.16), after evaluation of integrals in (2.9), take the form:

(2.18)
128 M. KACHANOV, ET AL.

Here are the incomplete

elliptic integrals of the first and the second kinds, respectively, and = arcsin(1/ ).
We now derive the expression for the stress vector t on the surface of the ellipsoid
= 0 . We express functions B m in the alternative form, namely, as potentials of a
simple layer, with t being the density of the layer:

(2.19)

On the other hand, relations (2.6), (2.8) and (2.13) yield the following expressions for
components of vector B:

(2.20)

The theorem on discontinuity of the normal derivative of the potential of a simple layer
implies that

(2.21)

Therefore, based on (2.20), we have:

(2.22)

Further, utilizing (2.11), (2.14) and the third relationship of (2.17), the tractions on the
ellipsoids surface = 0 are obtained as follows:

(2.23)

0
Components of the resultant force T that is required to produce translation u
(obtained by integration of (2.23) over the ellipsoids surface) are given by the following
stiffness relations:
ELASTIC SPACE WITH AN INCLUSION 129

(2.24)

The resultant moment M = 0 .

2.2. FIELDS PRODUCED BY ROTATION OF AN ELLIPSOID

The solution due to an arbitrary rotation of the ellipsoid is obtained by taking


harmonic vector B and harmonic scalar B 0 in Papkovich-Neubers representation (2.5)
in the form:

(2.25)

Here, functions 1 ( ), 2 ( ) and 3 () are given by (2.9) and 4 ( ), 5 ( ) and


6 ( ) are as follows:

(2.26)

Nine constants N m , D m , D' m (m = 1,2,3) are to be determined from the boundary


condition u = 0 = R 0 , or, in components:

(2.27)

Substitution of (2.25) into (2.5) and utilization of (2.11) yield expressions for the
displacement components. Boundary conditions (2.27) lead to a cumbersome system of
nine linear algebraic equations for the constants N m , D m , D ' m . However, solving this
system can be simplified by utilizing the superposition: the problem for an arbitrary
rotation vector is represented as system of the three sub-problems corresponding to
components 1 , 2 , 3 . Each of these sub-problems gives rise to a system of only three
equations for three constants.
In the case of 2 = 3 = 0:
130 M. KACHANOV, ET AL.

(2.28)

In the case of 1 = 3 = 0:

(2.29)

In the case of 1 = 2 = 0 :

(2.30)

Finding the constants this way yields the displacement field due to an arbitrary rotation
(outside of the ellipsoid, > 0 ) as follows:

(2.31)

where the following notations are used:

(2.32)
ELASTIC SPACE WITH AN INCLUSION 131

and where i ( ), after evaluation of integrals in (2.26), takes the form:

(2.33)

Functions (m = 1,2,3) are obtained from g m ( ) by changing the first term of


g m () according to the following rule:

Note that
Traction t on the ellipsoids surface is obtained by using the theorem on
discontinuity of the normal derivative of the potential of a simple layer, see (2.21) and
utilizing (2.11), (2.14) and the third relationship of (2.17). The expression for t is as
follows:

(2.34)
132 M. KACHANOV, ET AL.

Components of the resultant moment M that is required to produce rotation


(stiffness relations) are:

(2.35)
where

(2.36)

The resultant force T = 0 .

3. Oblate Spheroid

In this case, a 1 = a2 ( a 0 ) . The solution in this case cannot be obtained from the one
for the general ellipsoid by a straightforward substitution, but requires a non-trivial
limiting procedure. This is related to the fact that the ellipsoidal coordinates have to be
changed to the oblate spheroidal ones.
In all the equations for the general ellipsoid, we impose the following condition:
e 0, v 0 in such a way that the ratio v /e remains finite. We also set:

(3.1)

The relationships between cartesian coordinates (x1 ,x 2 , x3 ) and curvilinear coordinates


( s , q , ), obtained from (2.4) by the limiting transition, take the form:
ELASTIC SPACE WITH AN INCLUSION 133

(3.2)

where 0 s < , 1 q 1. The boundary of the oblate spheroid corresponds to s = s0


and is given by the following equation in Cartesian coordinates:

(3.3)

3.1. FIELDS PRODUCED BY TRANSLATION OF AN OBLATE SPHEROID

Utilizing (3.1-3), the expressions in (2.17,18) can be reduced to elementary functions, as


follows:

(3.4)

The displacement field outside of the spheroid, s > s0 , due to translation u0 takes the
form:

(3.5)

where the following notations are used:

(3.6)
134 M. KACHANOV, ET AL.

Traction vector t on spheroids surface s = s 0 is:

(3.7)

Components of resultant force T , that is required to produce translation u 0 , are


given by the following stiffness relations:

(3.8)

The resultant moment M = 0 .

3.2. FIELDS PRODUCED BY ROTATION OF AN OBLATE SPHEROID

The expressions in (2.33), in the case of the oblate spheroid, can be reduced to the
following elementary functions:

(3.9)
ELASTIC SPACE WITH AN INCLUSION 135

The displacement field outside of the spheroid, s > s 0 , due to rotation , after some
algebra, takes the form:

(3.10)

where the following notations are used:

(3.11)

Traction vector t on the oblate spheroids surface, s = s 0 , due to rotation , is:

(3.12)

Components of resultant moment M , that is required to produce rotation , are


given by the following stiffness relations:
136 M. KACHANOV, ET AL.

(3.13)

where

The resultant force T = 0 .

4. Prolate Spheroid

In this case, a 2 = a 3 Similarly to the case of the oblate spheroid, the

solution cannot be obtained from the one for the general ellipsoid by a straightforward
substitution, but requires a non-trivial limiting procedure. This is related to the fact that
the ellipsoidal coordinates have to be changed to the prolate spheroidal ones.
In all the equations for the general ellipsoid, we impose the following conditions:

(4.1)

The relationships between cartesian coordinates (x 1 , x 2 , x 3 ) and curvilinear coordinates


( s, q, ), obtained from (2.4) by the limiting transition, take the form:

(4.2)

where 1 s < , 1 q 1. The boundary of the prolate spheroid corresponds to s = s 0


and is given by the following equation in Cartesian coordinates:

(4.3)
ELASTIC SPACE WITH AN INCLUSION 137

4.1. FIELDS PRODUCED BY TRANSLATION OF A PROLATE SPHEROID

Utilizing (4.1-3), the expressions in (2.17,18) can be reduced to elementary functions, as


follows:

(4.4)

The displacement field outside of the spheroid, s > s 0 , due to translation u 0 , takes the
form:

(4.5)

where the following notations are used:

(4.6)

Traction vector t on the spheroid's surface s = s 0 is:


138 M. KACHANOV, ET AL.

(4.7)

Components of resultant force T , that is required to produce translation u0 , are


given by the following stiffness relations:

(4.8)

The resultant moment M = 0 .

4.2. FIELDS PRODUCED BY ROTATION OF A PROLATE SPHEROID

The expressions in (2.33), in the case of the prolate spheroid, can be reduced to the
following elementary functions:

(4.9)

The displacement field outside of the spheroid, s > s0 , due to rotation , after some
algebra, takes the form:

(4.10)
ELASTIC SPACE WITH AN INCLUSION 139

where the following notations are used:

(4.11)

Traction vector t on the prolate spheroids surface, s = s 0 , due to rotation , is:

(4.12)

Components of resultant moment M , that is required to produce rotation , are


given by the following stiffness relations:

(4.13)
140 M. KACHANOV, ET AL.

where

The resultant force T = 0 .

5. Rigid Sphere

We now consider the simplest case of a rigid sphere of radius R 0 embedded into an
infinite elastic medium. It is given small translation u 0 and small rotation , so that
displacements of the points of the spheres boundary are:

(5.1)

where R 0 is the position vector of points on sphere's surface.


To determine the displacement and stress fields outside of the sphere, we utilize
Trefftz's general solution for displacements. The displacement field, outside of the sphere
( R > R 0 ), can be obtained in the following form:

(5.2)

(5.3)

In the text to follow, we transform these expressions to a more convenient representation


of displacements (and stresses) in spherical coordinates (R , , ), with unit vectors
e R , e , e .

5.1. FIELDS PRODUCED BY TRANSLATION OF A SPHERE

In this case u 0 = u 0 i , where i is the unit vector along x - axis. Using the relationship

(5.4)

where H R =1, H = R , H = R sin are Lames coefficients, we have:

(5.5)
ELASTIC SPACE WITH AN INCLUSION 141

and, since R / R = e R , we also have

(5.6)

After substitution of (5.5,6) into (5.2) and some algebra the displacement components in
spherical coordinates take the form:

(5.7)

and the dilatation is:

(5.8)

Stress components in spherical coordinates are obtained as follows:

(5.9)

On the surface of the sphere ( R = R 0 ), the stresses are:


142 M. KACHANOV, ET AL.

(5.10)

so that the traction vector on the sphere's surface is:

(5.11)

The resultant force T, required to produce translation u 0 , is given by the following


stiffness relation:

(5.12)

The resultant moment M = 0.

5.2. FIELDS PRODUCED BY ROTATION OF A SPHERE

We now consider the displacement field u due to rotation vector = 0 k. Since


k = e R cos e sin , substitution of into (5.3) gives the only non-zero displacement
component in spherical coordinates as follows:

(5.13)

The only non-zero stress component is:

(5.14)

The traction vector on the surface of the sphere is:

(5.15)

The resultant moment M, required to produce rotation , is given by the following


stiffness relation:
ELASTIC SPACE WITH AN INCLUSION 143

(5.16)

The resultant force T = 0 .

6. Conclusions

Closed form expressions are derived for the full set of elastic fields generated by a rigid
ellipsoidal inclusion embedded into an infinite elastic space and subjected to (arbitrarily
oriented) translations and rotations. Stiffness relations that interrelate the displacements
and rotations of the inclusion to the forces and moments applied to it are also given.

Acknowledgement. The authors are grateful to Dr. Rahman for making a preprint of his
work available. This work was supported by the National Science Foundation and U.S.
Department of Energy through grants to Tufts University.

References

Eshelby, E.J. (1961) Elastic inclusions and inhomogeneities. Progress in Solid Mechanics, eds. Sneddon J.N.,
Hill R., North-Holland, Amsterdam, V. 2, pp. 89-140.
Kanwal R.P and Sharma D.L. (1976) Singularity methods for elastostatics. J. Elasticity, 6, pp.405-418
Keer, L.M. (1965) A note on the solution for two asymmetric boundary value problems. Int. J. Solids
Structures, 1, pp. 257-264.
Lure, A.I. (1970) Theory of Elasticity. Nauka, Moscow (in Russian).
Rahman, M. (2000) The normal shift of a rigid elliptical disk in a transversely isotropic solid. Int. J. Solids
Structures (in press).
Selvadurai A.P.S. (1976) The load-deflection characteristics of a deep rigid anchor in an elastic medium.
Geotechnique, 26, pp. 603-612.
Selvadurai A.P.S. (1979) On the displacement of a penny-shaped rigid inclusion embedded in a transversely
isotropic elastic medium. SM Arch., 4, pp. 163-172.
Selvadurai A.P.S. (1980) Asymmetric displacements of a rigid disc inclusion embedded in a transversely
isotropic elastic medium of infinite extent. Int. J. Engng. Sci., 18, pp. 979-986.
Zureick A.H. (1988) Transversely isotropic medium with a rigid spheroidal inclusion under an axial pull.
ASME J. Appl. Mech., 55, pp. 495-497.
Zureick A.H. (1989) The asymmetric displacement of a rigid spheroidal inclusion embedded in a transversely
isotropic medium. Acta Mech., 77, pp 101-110.
This page intentionally left blank.
STRAIN PERCOLATION IN METAL DEFORMATION

R. M. THOMSON
Emeritus, Materials Science Engineering Laboratory,
NIST, Gaithersburg, MD 20899
L. E. LEVINE
Materials Science Engineering Laboratory,
NIST, Gaithersburg, MD 20899
AND
Y. SHIM
Center for Simulational Physics,
University of Georgia, Athens, GA 30602

Abstract. In previous papers, we have introduced a percolation model


for the transport of strain through a deforming metal. In this paper, we
summarize the results from that model, and discuss how the model can be
applied to the deformation problem. In particular, we outline the primary
experimental features of deformation which the model must address, and
discuss how the model is to be used in such a program. It is proposed
that the discrete percolation events correspond to slip line formation in a
deforming metal, and it is shown that the deforming solid is a self organizing
system. It is recognized that deformation is localized in space and time, that
deformation is fundamentally rate dependent, that hardening depends upon
relaxation processes associated with discrete percolating events, and that
secondary slip is an essential part of band growth and relaxation processes.

1. Preface

It is a pleasure to participate in this celebratory volume to Jim Rice. Our


collaboration on the problem of dislocation emission from cracks began
in the Summers at the ARPA Materials Research Council, and eventually
matured in our 1974 paper. I learned from that experience that Jim is one
of those rare people who can work creatively in more than one field at once,
145
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 145157.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
146 R. M. THOMSON ET AL.

because he so quickly grasps the essence of the main issues, even when the
scientific language and usage is quite different from his own. In addition,
I have been privileged to be a friend of an unusually kindly and generous
person over the years.
Jim, I wish you many more years of productive and distinguished ac-
complishment in science!

2. Introduction

A successful theory of work hardening in metals would include both an


analysis of the pattern formation of dislocations during metal deformation
and the transport of mobile dislocations through the dislocation structures
that are formed. A treatment of pattern formation would include the change
from one type of pattern to another (e.g. carpets to cells) as well as the
evolution of the size, distribution and shape of a basic pattern as deforma-
tion proceeds. We have not attempted such a complete theory, and suspect
that a satisfactory attempt is still some distance in the future. Rather, we
have noted that the pattern formation and transport parts of the problem
are logically rather different, and have chosen to focus on just the trans-
port half. In our work, then, the structure and its evolution is an input,
although we find that certain features of the transport problem are useful
in understanding the pattern size evolution. A useful dividend of our focus
on the transport problem is that the constitutive laws and stress/strain
relations are closely related to the transport problem, and should follow
rather straight forwardly from an adequate understanding of it.
Another important part of our approach and our motivation is the re-
alization that the metal deformation problem is dominated by probability
and statistics, and we have been guided by a hope that developments in
modern statistical physics and critical phenomena might provide useful in-
sights and concepts for understanding the deformation problem. We believe
this work has been at least a partial success in implementing that hope.
One of the most active areas in modern research on deformation has
been the direct application of computer simulation to deformation, and
the development of what is now called dislocation dynamics. We believe
that the work here is complementary to dislocation dynamics in the sense
that our model is viewed as a statistical theoretical framework, not a first
principles theory, and contains a variety of internal variables and input
parameters, which must be obtained from outside the model. These inputs
are expected to be supplied by some combination of incisive experiment
and dislocation dynamics addressing the question of how dislocation cells
operate both as sources and sinks for mobile dislocations.
This paper will first provide an overview of the model and its primary
STRAIN PERCOLATION IN METAL DEFORMATION 147

mathematical features and results, including a universal stress/strain re-


lation. We then summarize the essential experimental findings to which
the model must relate, discuss time dependent effects, and finally, list ex-
perimental and dislocation dynamics studies that are needed to provide
essential inputs to the overall modeling effort.

3. The Model
The model ((Thomson and Levine, 1998; Thomson et al., 2000)) is built
on the premise that the system has been brought to a level of strain cor-
responding roughly to stage III, where a cellular structure has developed.
(The model can probably be applied to more general situations, but these
have not yet been explored.) The basic idea is that the walls of the cell
structure provide both sources of mobile dislocations which move through
the interior of the cell, and barriers to the fully free motion of those dis-
locations. When mobile dislocations are arrested by the other walls of the
cell in which they have been formed, they exert concentrated stresses on
the incipient sources in these arresting walls, and the process repeats. In
detail:
1. The system is brought to a definite state of unidirectional strain char-
acterized by the plastic strain, p , and a flow stress,
2. The weakest cell in a band of slip undergoes a burst of strain which
initiates a cluster of newly strained cells. The initiating strain is s0 .
For convenience, we will simply take the strain variable for each cell as
the number of mobile dislocations created in that cell during a specific
percolation event.
3. The transmission of strain from a strained cell to a neighboring un-
strained cell is given by
s* = as, (1)
where s is the number of dislocations in the strained cell, s* is the
number induced in the previously unstrained neighboring cell, and a is
the amplification factor, a stochastic function characterizing the wall
between a strained and unstrained cell.
4. Because s corresponds to the discrete number of mobile dislocations in
a cell, s* 1. (We do not take the trouble to discretize the variable,
but it is necessary that it always be at least unity.)
5. Growth takes place only on the boundary of a strained 2-d cluster of
cells, because any cell strained in a percolation event will be hardened
against further straining during that event.
6. The stochastic amplification factor depends on the properties of the
wall, and can be written a = a ( P1, P 2 ,). where P 1 , P 2 , etc. are pa-
rameters. We visualize at least two separate mechanisms for transmit-
148 R. M. THOMSON ET AL.

ting strain from cell to cell, a source mechanism and a lock breaking
mechanism. We have investigated two possible cases. In case I, only
the source mechanism operates. In case II, both mechanisms play a
role.
In the source mechanism, the wall possesses a distribution of disloca-
tion sources of varying strength whose maximum strength is given by
the parameter P1 . For case I, we shall write

(2)

where is a random number.


In the second mechanism for transmitting strain, we presume that
the cell walls are localized in the lattice by locks of various sorts and
strengths, such as Lomer-Cottrell locks, and that these locks can be
broken, (unzipped) under the stress of the pile-up dislocations. When
that happens, the region of the wall stabilized by the lock will break
away as mobile dislocations into the next cell, with a large amplification
factor. For case II, we then modify Eqn. (2) to give

(3)

where P 2 is the amount of strain released in an unzipping event and


k is a measure of the probability of such an event occurring. It will
be assumed that P 2 >> 1 and k << 1 ,corresponding to the assumption
that unzippable locks are relatively rare, but when they do occur, the
resulting strain is large.
7. In a computer implementation, one begins a simulation at the cell at
the origin, for which an initiating value of strain, s 0 , is assumed. Given
the rules for strain transmission, the initiation value must be greater
than a limiting critical value, s 0 > 1/P 1 , in order for the strain to grow
out of the origin. (This critical value is only correct if, as in case I, the
sites next to the origin have no walls with unzippable locks.)
These rules constitute a well posed percolation problem, and a unique per-
colation threshold is observed at P1 = P1 c in case I. In case II, a percolation
threshold also exists, but in this case the threshold lies on a critical sur-
face, C ( P1 , P 2, k ) = 0. We have shown that the geometrical aspects of the
percolating strain clusters conform to the universality class of ordinary per-
colation theory ((Stauffer and Aharony, 1992)). Results for case I have been
published ((Thomson et al., 2000)). Fig. 1 shows a typical run for case II
for a cluster which spans the system size chosen at the critical point.
STRAIN PERCOLATION IN METAL DEFORMATION 149

Figure 1. Spanning cluster at the critical point. Dark regions represent larger strain
and lighter represent smaller strain in a cell. Here, s 0 = 2.2, P 2 = 40, and k 0 = 0.01 for
L = 401 with P 1c = 0 . 6 4 5 2 . L is the linear dimension of the system.

Since the strain percolation problem contains the strain variable, s ,


which has no counterpart in standard percolation theory, the scaling laws
for s must also be worked out. We have done so for case I ((Shim et al.,
2000)), but case II is still incomplete. In case I, the scaling law for the strain
is found to be
(4)

where T is the strain per cell site in the system, = 2.389 is one of the ex-
ponents of standard percolation theory, S is the average number of strained
sites in the system, and is a new exponent whose value is determined to
150 R. M. THOMSON ET AL.

be = 0.75 0.25 (one standard deviation), and c is a constant. The ratio


Tc /Sc is the strain per strained site in the system at the critical point. For
this ratio to be finite, it is necessary that T have the same fractal dimension
at the critical point as does the geometric cluster, with
d = 1.896.
In the discussion immediately following, the behavior of the strain above
the percolation threshold is found to be the important physical quantity.
Equation (5) is valid both below and above the critical point, but above
the critical point, the law takes the simpler form,

(5)

where = 0.14 is one of the standard percolation parameters.


Since the geometrical density of strained sites per system site grows as
above the critical point in case I (with just one perco-
lation critical parameter), the meaning of Eqn. (5) is that the strain growth
is controlled and dominated by the geometry above the critical point. This
super critical growth does not extend very far above the critical point, be-
cause as P1 P c increases, the system reaches a second critical point where
the average strain diverges in a strain avalanche. That is, immediately above
the critical point, even though the geometry has exceeded its percolation
threshold, and its fractal character has disappeared, the strain seems only
subliminally aware that something significant has happened. That is, if we
examine the strain per strained site, in the sys-
tem, has a well defined fixed value for large system sizes below
the critical point, and grows only slowly in nearly linear fashion above the
critical point, increasing only about 40% in the super critical region below
the avalanche. Since this contribution to the strain per cell is only slowly
varying in the super critical regime, and the geometrical density of strained
sites is increasing quite rapidly, the geometrical increase dominates. That
is, when one plots the strain per cell in the super critical regime, the linear
increase mentioned above gets lost in the strong increase which signals the
approach of the avalanche. This change in ( P ) can therefore be realistically
neglected when plotting T ( P ), below the avalanche, so that T can simply be
regarded as exhibiting the geometrical increase of strained clusters above
the percolation critical point. But it is interesting that the mathemati-
cal model exhibits two separate critical points slightly separated from one
another: a percolation threshold and an avalanche point. Of course, any
physical significance of the avalanche point must be considered in the light
of the very severe limits one must put on the upper limit to the number of
dislocations which a wall can produce.
STRAIN PERCOLATION IN METAL DEFORMATION 151

4 . A Universal Constitutive Law.

In the following, we are mainly interested in generic results, so we shall


limit ourselves to case I with the single percolation parameter, P1 .
We postulate that the model corresponds to actual strain percolation
events in a real system, and that the strain in the deforming solid is the
accumulated strain in the percolation events. It turns out that the strain at
the percolation threshold is exactly zero in the thermodynamic limit of
infinite system size, and thus finite strain only appears slightly above the
threshold in the super critical regime. Thus, as the system moves along the
stress strain curve, it remains always at the critical threshold, or actually
just slightly above it. This behavior defines a self organizing critical (SOC)
system.
Since the percolation parameters describe the production of dislocations
from the cell walls, and since this ability depends on both the flow stress,
as well as the state of strain in this unidirectional straining system, the
percolation parameters must depend on the strain, , and the stress, .
Thus, in the simple case I, P1 = P 1 ( , ), and on the stress/strain curve,

(6)

(A somewhat more complicated, but equivalent, treatment follows for case


II, where the weak walls are included.) If one remembers that the perco-
lation events are discrete, then from one percolation event to the next, we
can write the stress/strain law as a universal relation,

(7)

This equation is rigorous, and follows from the SOC of the deformation,
which is the reason we call it a universal law. But it can take on physical
meaning only when it is possible to relate the percolation parameter to the
stress and strain. In . view of the way the model is constructed, and its fun-
damental relation to the way the cell walls create and obstruct mobile dis-
locations in the system, this functional dependence can only be determined
by a detailed study of what we term dislocation cell physics. Hopefully,
from this study, the two physical mechanisms by which we believe dislo-
cations are produced (source action and unzipping) can be distinguished,
and stochastic distributions determined, from which percolation variables
can be deduced.
It is in this sense that we claim that the strain percolation model pro-
vides a stochastic framework for a deformation theory, but the inputs to
this framework must be determined by separate study of the dislocation
cell phenomenology.
152 R. M. THOMSON ET AL.

In the following section, we will take some tentative steps in this direc-
tion by looking at what the existing experimental lore can tell us about the
cell physics, and ultimately the constitutive relations.

5. Real Slip

In this section, we review the major features of slip in a deforming metal to


which the percolation model must conform. The model is already consonant
with the existence of the cell structure, by its nature. But slip in metals is
localized in space and time in characteristic ways which we now explore.
In all of the standard stages of strain, the elementary slip events are
observed as fine slip lines on the surface of the deforming metal. Up through
Stage III, these lines are many micrometers in length (but they shorten
with strain), and are up to several hundred Burgers vectors in height. We
will identify the slip which occurs in a percolation event in our percolation
model with this observed universal elementary unit of slip.
A striking observation which begins in Stage II, and is fully developed in
Stage III, is the spatial localization of the percolation events into bands of
slip. This localized ordering of the percolation events into bands is critical
to understanding the stress-strain law, because the percolation events in a
band interact strongly with one another.
Our principal focus is on aluminum, and the quantitative study of the
bands in Al goes back to early electron microscope replica studies by H.
and D. Wilsdorf ((Wilsdorf and Kuhlmann-Wilsdorf, 1952)) and by Noggle
and Koehler ((Noggle and Koehler, 1957)). In addition to this early work,
we are aware of a more modern measurement of Al band structure by W.
Tong et al. ((Tong et al., 1997)) using atomic force microscopy (AFM). One
of the authors, H. Weiland ((Weiland, )), reported some additional AFM
results to us privately. These AFM measurements are consistent with the
early and much lower resolution electron microscope replica work.
All of the results are quoted for room temperature aluminum at strain
levels where the cell structure should be well developed. Weiland differs
from the other authors in that his aluminum samples were doped with Mg,
while the others worked with high purity metal.
A summary of the main conclusions of these studies is:
1. About 80% of the total strain is concentrated in the bands. But the
minority strain taking place in the matrix means that the matrix is
hardening along with the bands.
2. Both bands and slip lines are very long compared to a typical cell size.
3. The number of slip lines contained in a band varied from a few to the
order of 100, with an average in the range of 20. This number increased
slowly with total strain.
STRAIN PERCOLATION IN METAL DEFORMATION 153

4. The slip line height increases with strain, but in general, the authors
find heights between about 10 and several hundred Burgers vectors,
with an average in the neighborhood of 30-50 ((Noggle and Koehler,
1957; Weiland, )).
5. The distance between slip lines appears to have a minimum value of
about 30 atomic distances, so there is a natural width to a percolation
event.
Data on the time localization of slip are available from experiments
reported by Pond ((Pond Sr., 1972)) on high purity Al at room temperature,
but total strain levels are not reported. From the authors comment that
measurements were made till the slip lines interfered with one another,
it is presumed that these data are also representative of strain at a level
where the cells were well developed. The data were obtained from optical
cinematography on the metal surfaces.
Pond finds that a slip band grows in discrete jumps, with individual
jumps taking place in about 0.1 sec., and with several to many seconds
between jumps. The amount of growth in a jump varies from 50 b to 500
b, with an average around 150 b to 200 b. If these data can be correlated
with the slip line heights reported above, it appears that the band growth
takes place with the production of a small number (around 3 to 7) of slip
lines in an average growth event.
In summary, the slip is localized in both time and space, probably with
multiple percolation events energizing one another in the discrete growth
jumps occurring in a band. The band grows by both filling in the allowable
space between slip lines (percolation events) and by growth in width of the
band. (In experiments in Cu ((Mader, 1957)), it is known that individual
bands broaden with strain, and that new slip bands are nucleated out of
the matrix, but corresponding experiments have not been done in Al.) The
slip line heights are greater at larger strain, which may be because multiple
percolation events can occur in the same location, or perhaps because the
slip generated in a percolation event at the higher strain levels is greater.
Clearly, the phenomenological picture needs much clarification, especially
as regards the quantitative aspects of band growth.

6. Relaxation Modeling of Bands

We have shown how a universal stress-strain law follows from a simple


assumption that our discrete percolation model contributes all the strain in
the deforming metal. If the percolation events of our model are identified
with the elementary slip lines observed in all deforming metals, one can
have high confidence that it is correct. But there is much physics hidden in
the universal law, and we will now construct an additional level of detail
154 R. M. THOMSON ET AL.

based on the findings we have summarized in the previous section. In the


current section, we focus on the growth of a well developed band containing
a relatively large number of elementary slip lines.
The first question that immediately asserts itself is why the bands exist
in the first place? Equivalently, why is a local region where multiple slip lines
have been produced softer than the matrix? There is not space here to enter
into an adequate discussion of this question, but we believe the relevant
mechanism is the rotation created by the, not necessarily local, interaction
between the majority primary dislocations which have been created in the
slip lines, and the secondary dislocations.
In brief, the secondary dislocations which are produced in the slip band
are all of such a character as to rotate the lattice in a slip band region in
a softening direction relative to the matrix. That is, the rotation is such as
to increase the resolved shear stress on the primaries. Since at the start of
a percolation event, the flow stress in the band region must be the same
as that in the matrix, because the matrix and band regions both deform,
the rotation of the band constitutes an instability of the band relative to
the matrix. This mechanism has its seat in the uncompensated dislocations
which are produced in regions of strain gradients.
We now consider in a very generic way the effect of a percolation event
on the stress level in a band. In the following, we shall not distinguish
between a single percolation event and several correlated simultaneous per-
colation events. If we assume the system is in a hard tensile machine with
a prescribed strain rate, then the strain, , in the sample can be written as
a linear function of time, t, as

(8)

where subscripts p and e refer to the plastic and elastic parts. We consider
the time between one (the reference) percolation event and the next. At
the moment of the event when t = 0, the strain in the system is increased
suddenly by the strain, p , in the percolation event, and the stress in the
system is lowered by the amount, p , where is the elastic constant.
After time zero, the stress increases linearly because of the way the system
is loaded, and (t) = ( t p ). When the actual stress, (t), reaches the
flow stress, p (t), another event will occur, and the system cycles.
By definition, the system at time t = 0 was at the flow stress for that
state of strain, and if the flow stress does not change, then the time be-
tween events will be given by the time for the tensile machine to build the
stress back to the initial level. But the flow stress is altered by the strain
in the percolation event. The flow stress in the slip plane of the event(s)
is immediately increased because of the local increase of immobilized dis-
locations in the slip line plane(s). In the time provided by the stress drop
STRAIN PERCOLATION IN METAL DEFORMATION 155

and subsequent build up, the immobilized dislocations will recover by cross
slip, climb, etc., thereby altering the flow stress both within the immediate
slip plane of the event as well as a region parallel to that plane which is
reachable by short range cross slip and climb processes.
In addition, secondary slip will be initiated by the percolating primary
slip, and these non-percolating secondary dislocations will move on slip
planes at a large angle to the primary plane. Presumably the secondaries
have a mean free path of the order of a single cell size. Their effect is three
fold. First, they will contribute forest dislocations to cell walls throughout
the active band, and second, they will interact with incipient sources on
other slip-line planes in the band. This second interaction is a mechanism
for initiating multiple percolation events simultaneously, and it is also a
possible mechanism for enhancing the source distribution throughout the
band for a subsequent percolation event. Third, the additional secondaries,
in their interaction with the primary dislocations of the band will rotate
the band relative to the matrix, thereby softening it.
This complex set of processes has the effect of increasing the flow stress
of the band over the time between successive slip events, but its precise
form is very difficult to surmise without more detailed knowledge about
the cell physics. We write this unknown function as

(9)

because the stress relaxation function, R(t), must be linear in the strain
increment of the percolation event. ' is a constant. The next event will
occur at the time when the new flow stress is equal to the stress building
in the deforming sample,

(10)

which has a complicated implicit dependence on the time. However, since


the slope of the stress strain curve in the plastic regime is much smaller
than the elastic constant, , we can take the time between events as simply
t = p / , and the stress-stain relation follows,

(11)

In writing this equation, we make the further assumption that R(t) is a


slowly varying function in the neighborhood of t = t . This equation now
fleshes out some of the underlying relaxation physics in the earlier universal
relation, by focusing attention on the mechanisms by which the flow stress
relaxes during band growth. It also emphasizes that deformation in metals
in Stage III is at its most fundamental level, a rate process. Finally, it shows
that deformation at this most fundamental level is a discrete process, and
156 R. M. THOMSON ET AL.

that the average strain in a discrete percolation event, p , is a fundamental


quantity, which appears explicitly in the final stress-strain relations. The
connection between the universal equation and the relaxation equation ap-
pears when one relates the percolation parameter, P 1 c to the relaxing flow
stress, p .

7. Conclusions and Needed Experiments

The universal stress-strain law obtained from the SOC character of the de-
forming metal must be extended to include the consequences of the ordering
of the percolation events into spatially localized slip bands. This ordering
points to the importance of the relaxation of the flow stress between time-
localized clusters of percolation events in the band.
We believe the identification of mechanisms for band growth and the
relaxation of the flow stress between time-localized clusters of percolation
events requires new experiments and modeling focused on these issues.
Specifically:
1. Use AFM to obtain 3D plots of the slip-line and slip-band structures
and explore the growth of the bands, hopefully as a function of orien-
tation, alloy composition, strain and strain rate.
2. Correlate the AFM measurements with the underlying cell structure
obtained by TEM and synchrotron X-ray measurements.
3. Do time localization studies of band growth bursts by acoustic emis-
sion, hopefully with correlation of the time bursts with particular
bands.
4 . Model cell wall structure and source formation by dislocation dy-
namics. If computationally possible, also explore band formation and
growth by dislocation dynamics.
Results from such experiments and modeling will supply information
on the parameters involved in the model, and also on the strain, , of an
individual percolation event. This quantity is an output of our model, but
we have not studied how case II behaves above the critical point sufficiently,
and we believe that is the relevant physical case. Further, the observed
values of in Ponds experiments seem to be of the order 50, which is rather
large compared to our expectations for wall source maximum outputs. But
Ponds experiments may involve more than one percolation event in a strain
burst, as discussed in the text, so the connection between the experiments
and the theory must still be considered uncertain. Thus, further study is
required both of the model and experimental measurements of this critical
parameter, before we can present a satisfactory physical picture.
Finally, we note that the percolation model has implicit in it a mecha-
nism for cell evolution through the nucleation of new cell walls by means
STRAIN PERCOLATION IN METAL DEFORMATION 157

of the capture of mobile dislocations in mid-cell by secondary dislocations.


Incorporation of such cell evolution features is required for producing a
complete stress-strain law.

References
Mader, S. (1957). Z. Phys., 149:73.
Noggle, T. and Koehler, J. (1957). J. Appl. Phys., 28:53.
Pond Sr., R. B. (1972). The inhomogeneity of plastic deformation. In Reed-Hill,
R. E., editor, ASM Seminar Series, page 1, Metals Park, Oh. ASM.
Shim, Y., Levine, L., and Thomson, R. (2000). Mater. Sci. Eng. A. in press.
Stauffer, D. and Aharony, A. (1992). Introduction to Percolation Theory. Taylor
and Francis.
Thomson, R. and Levine, L. E. (1998). Theory of strain percolation in metals.
Phys. Rev. Lettr, 81:38843887.
Thomson, R., Levine, L. E., and Stauffer, D. (2000). In press.
Tong, W., Hector, Jr., L. G., Weiland, H., and Wieserman, L. F. (1997). In-situ
surface characterization of a binary aluminum alloy during tensile deformation.
Scripta Mater., 36(11):13391344.
Weiland, H. private communication.
Wilsdorf, H. and Kuhlmann-Wilsdorf, D. (1952). Z. Ang. Phys, 4:23.
This page intentionally left blank.
DIFFUSIVE INSTABILITIES IN DILATING AND
COMPACTING GEOMATERIALS
J. W. RUDNICKI
Department of Civil Engineering
Northwestern University
2145 Sheridan Road
Evanston, Illinois 60208-3109

Abstract. This chapter reviews and extends analyses of diffusive instabilities in


inelastically deforming geomaterials. The onset of these instabilities is connected
with the conditions for shear localization in the limiting cases of drained (constant
pore pressure) and undrained (constant fluid mass) deformation and depends on
whether inelastic volume change is dilation or compaction. Rice [1975] showed that
homogeneous shear deformation of a layer was stiffer for undrained than for drained
conditions but was unstable in the sense that the magnitude of infinitesimal spa-
tial nonuniformities begins to grow exponentially in time when the condition for
localization is met in terms of the underlying drained response. As the condition
for localization in terms of the undrained response is passed, infinitesimal spatial
perturbations experience infinitely rapid decay and then infinitely rapidly growth.
For materials that dilate with inelastic shearing the condition for localization is
met for the drained response before it is met for the undrained response. For ma-
terials that compact and for which the shear yield stress increases with normal
stress, the undrained response is softer than the drained and conditions for local-
ization are met for undrained response before drained. If the shear yield stress
decreases with normal stress, as for materials modeled by a cap on the yield
surface, results for compacting materials are identical to those for the dilating ma-
terials. Generalization of the layer results to arbitrary deformation states reveals
the same relation for the onset of diffusive instability: spatial nonuniformities be-
gin to grow exponentially when the condition for localization is met in terms of
the underlying drained response. In contrast to the result for the layer, the growth
rate of perturbations does not necessarily become unbounded when the condition
for localization is met in terms of the undrained response. The difference is due
to a lack of symmetry in the constitutive tensors that is typical of geomaterials.
Explicit expressions are given for the undrained response in terms of the drained
for an elastic-plastic relation with yield stress and flow potential depending on
159
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 159182.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
160 J. W. RUDNICKI

first and second stress invariants. For this relation and the limit of incompress-
ible solid constituents, the lack of symmetry just-mentioned disappears. If the
fluid constituent is also incompressible, the analysis confirms a result of Runesson
et al. [1996] that the undrained response is independent of mean stress and the
predicted direction of shear bands is 45 to the principal axes of stress.
1 Introduction
In contrast to metals, inelastic deformation of geomaterials typically involves vol-
ume change. In low porosity rocks, dilatancy (volume increase) can occur during
inelastic shearing under compressive mean stress because of local tensile microc-
racking at the tips of sliding fissures or at local property mismatches, and from
uplift in sliding over asperity contacts on fissure surfaces. In soils, dilatancy re-
sults from rearrangement of close-packed particles due to shearing. Compaction
in high porosity rocks can result from the collapse of pore structures due to shear-
ing or high mean stresses and, at very high mean stresses, from grain crushing.
Compaction of low density soils occurs when shearing causes a closer packed ar-
rangement of particles.
When the geomaterial is fluid-saturated, inelastic volume changes tend to cause
a change in pore fluid pressure. If the deformation is slow enough and drainage
from the boundaries is possible, alterations in pore pressure will be equilibrated
by fluid mass flow. In this drained limit, the pore pressure is constant. For
volume changes that occur without allowing drainage from material elements, the
pore pressure changes. This undrained limit can occur if volume changes occur
too rapidly (though still slow enough so that inertia is not significant) to allow
time for fluid mass flow or during homogeneous deformation if fluid flow from the
boundaries of the body is prevented.
Because the inelastic response of geomaterials is affected by the mean effective
stress, that is, the total compressive mean stress minus the pore pressure, alter-
ations in pore pressure will either inhibit or promote further inelastic straining.
Consequently, the inelastic response differs in the drained and undrained limits
and, for intermediate cases, is coupled to the diffusion of pore fluid. Conditions
for failure and, in particular, for localization of deformation depend on this cou-
pling.
In a seminal analysis, Rice [1975] examined the coupling of pore fluid diffusion
and inelastic response for combined shear and compression of a layer that exhibited
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 161

dilatant volume changes. He showed that dilatancy during homogeneous shearing


of the layer without allowing fluid drainage from the boundaries prevented caused
a reduction of pore fluid pressure and, thus, an increase in effective compressive
stress. This increase inhibited further inelastic deformation. But Rice [1975]
proceeded to show that homogeneous dilatantly strengthened response becomes
unstable when the condition for localization of deformation [Rudnicki and Rice,
1975] is met in terms of the underlying drained (constant pore pressure) response.
For the dilatant behavior and layer model considered in Rice [1975], this condition
occurs when the drained shear stress versus shear strain curve reaches a peak even
though the undrained response curve is still rising. When this condition is met,
spatial non-uniformities grow exponentially in time with the smallest wavelengths
growing the fastest. Because spatially nonuniform deformation causes fluid flow
in response to pore pressure gradients, homogeneous undrained response cannot
be realized beyond this point.
Dilatant strengthening has been widely observed in granular materials dating
back to the experiments of Reynolds [1885]. More recently, it has been observed
in laboratory tests on both rocks [Brace and Martin, 1968; Martin, 1980] and
soils[Mokni and Desrues, 1999]. Vardoulakis [1985, 1986, 1996a, b] has adapted
Rices analysis for the biaxial deformation of both dilating and compacting water
saturated sand and used it to interpret laboratory observations on the develop-
ment of localization of deformation. Recent theoretical analyses [Runesson et al.,
1996; 1998] have examined conditions for localization in the limit of undrained
deformation.
In this article, I review Rices [1975] analysis and discuss more generally its
implications, in particular, for compacting materials. Compaction can result not
only from inelastic shearing but also from purely hydrostatic stress. The inelastic
response of materials that compact under hydrostatic stress is often modeled by a
cap on the yield surface enclosing the hydrostatic axis [Dimaggio and Sandler,
1971; Wong et al., 1997]. The presence of this cap implies that inelastic shearing
is enhanced, rather than inhibited, by an increase in mean compressive stress.
Recently, Issen and Rudnicki [2000] have shown that for compacting materials
the conditions for localization admit solutions not only for shear bands, but also
for compaction bands. Compaction bands, narrow planar zones of compacted
material that form perpendicular to the largest compressive principal stress, have
been observed both in the field [Mollema and Antonellini, 1996] and in laboratory
experiments [Olsson, 1999].
For the deformation state considered in Rice [1975] the conditions for localiza-
tion are met at the peak of the shear stress versus shear strain curve. For more
162 J. W. RUDNICKI

general deformation states, Rudnicki and Rice [1975] have shown that the con-
dition for localization may be met before or after peak stress. Here, I show the
conclusions of Rice [1975] concerning the stability of undrained deformation can
be generalized to arbitrary deformation states. A previous analysis of this type
has been outlined in Rudnicki [1983] but simplifies the description of fluid flow.
Rudnicki [1983] assumes the pore pressure is uniform in a planar band and in
the surrounding material, and, in addition, that fluid mass exchange between the
layer and the surrounding material is proportional to the difference in pore fluid
pressures. These simplifications are avoided here by resolving the perturbations
from the uniform fields in terms of Fourier components relative to the putative
band. The implications of the results are examined for a general elastic plastic
model and used to illuminate the role of pore fluid compressibility on conditions
for localization in undrained deformation [Runesson et al., 1996].
The next section briefly summarizes Rices [1975] analysis. Succeeding sections
develop the extensions to compacting materials and arbitrary deformation states.
2 Rices Analysis

2.1 Formulation
The geometry of the problem considered by Rice [1975] is shown in Figure 1: plane
strain deformation of a layer extending indefinitely in the x-direction. Displace-
ments in the x and y directions are u (y, t) and v (y, t), respectively, where t is
time. Only the normal strain (y, t ) = v /y (positive in extension) and the shear
strain (y, t ) = u / y are nonzero. Stresses work-conjugate to and are normal
stress (positive in compression) and shear stress . Equilibrium (in the absence
of body forces) requires that and be uniform. Hence, they are functions only
of time. Other reaction stresses exist to maintain the constraints of zero strain in
the x and out-of-plane directions
Constitutive relations relate increments of and to increments of and .
For constant pore pressure, Rice [1975] gives these as follows:

(1a)

(1b)

The first term in each expression is the elastic portion of the increment; G and M
are elastic moduli. The second terms in (1) are the inelastic portions. For constant
, the hardening modulus H is related to H tan , the slope of a curve of versus ,
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 163

Figure 1: Geometry of the layer problem analyzed by Rice [1975].

by
(2)
Thus, H H tan, for H << G The yield surface is the boundary of the stress states
( , ) that cause only elastic response for a given state of inelastic deformation;
, the local slope of this surface, is referred to as a friction coefficient (Figure 2).
Thus, when H > 0, deformation increments tending to make d d are purely
elastic and the second terms in (1) are dropped. (When H < 0, elastic unloading
corresponds to d d .) Thus, increases in compressive normal stress inhibit
further inelastic deformation. Inelastic increments of volume strain d p are related
to inelastic increments of shear strain d p b y

(3)
where is a dilatancy factor.
When the pore pressure is not constant, the constitutive relations (1) are mod-
ified by replacing the increment of normal stress by an increment of the effective
normal stress, a linear combination of d and dp. Experiments on failure of rocks
(e.g., Paterson [1978]) suggest that p is the appropriate form for the effective
164 J. W. RUDNICKI

Figure 2: Geometric interpretation of the constitutive parameters used by Rice


[1975].

stress for inelastic response. In addition, Rice [1977] has argued on theoretical
grounds that this is the appropriate form for inelasticity due to microcracking
from the tips of sharp fissures and to frictional sliding on surfaces with small real
contact areas. Generally, a different form is needed for elastic deformation [Nur
and Byerlee, 1971; Rice and Cleary, 1976]. But, when the solid and fluid con-
stituents are much less compressible than the porous matrix, as for most soils,
p is also the appropriate form for elastic straining. In this case, equations (1)
become

(4a)

(4b)

An additional constitutive relation is Darcys law which, in the absence of body


forces, has the following form:

(5)
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 165

where q is fluid mass flow rate per unit area in the y direction and is the fluid mass
density. The coefficient is often expressed as the ratio of a permeability, with
dimensions of length squared (frequently measured in darcies; 1 darcy = 10 8 cm2 )
to the fluid viscosity. If the fluid phase is incompressible, the density is constant
and the equation expressing conservation of fluid mass is

(6)

Substituting (5) into (6) yields an equation relating gradients in pore pressure to
changes in volumetric strain:

(7)

2.2 Undrained Homogeneous Deformation


If drainage from the boundaries of the layer is prevented and the deformation
and properties are uniform, (5) yields q = 0 throughout the layer and, from (7),
increments in volume strain are zero. Setting d = 0 in (4) yields the following
expression for the change in effective normal stress:

(8)

Since M is an elastic modulus and positive, the pore pressure decreases for dilation
( > 0) at fixed normal stress. Substitution of (8) into (4a) reveals that the slope of
the shear stress versus shear strain curve (no longer at constant effective normal
stress) is still given by (2) but with the hardening modulus H replaced by the
augmented value:
(9)
2.3 Instability
Rice [1975] proceeds to show, however, that the homogeneous, undrained solution
is unstable with respect to small spatial perturbations in the strain or pore pres-
sure. In particular, linearization of the governing equations about the undrained
homogeneous solution yields the homogeneous diffusion equation for the pertur-
bations in pore pressure :

(10)
where the diffusivity c is given by

(11)
166 J. W. RUDNICKI

and the constitutive parameters are to be evaluated at homogeneous undrained


deformation. Perturbations with a Fourier wavelength grow exponentially at a
rate
r = 4 2 c /2 (12)
If both and are positive, then, since M > 0, H u n d r a i n e d > H and t he
homogeneous, undrained response is dilatantly hardened. But, because c ( 1 1 )
passes through zero from positive to negative when H = 0 , the magnitude of
spatial perturbations, instead of decaying exponentially, grow exponentially (12).
As noted by Rice [1975] this is analogous to running the heat equation backwards
in time: non-uniformities become more localized rather than more diffuse with
time. Thus, dilatantly hardened response becomes unstable when the underlying
drained response passes through a peak. Since H is generally a decreasing function
of inelastic deformation, H = 0, corresponding to a peak in the drained vs.
curve (at constant ) will occur before a peak in the undrained response curve,
H u n d r a i n e d = 0.
The Appendix of Rice [1975] develops the analysis for arbitrarily compressible
solid and fluid constituents. The effect is to replace the elastic modulus M in (8),
(9) and (11) by a modified value

(13)

where is the apparent void volume fraction, K is the bulk modulus of the pore
fluid and M s and N s are additional moduli associated with the solid constituents.
When both the solid and fluid constituents are effectively incompressible, M' = M.
If the solid constituents are incompressible, i.e., Ms , N s >> M ,

(14)

If the fluid is very compressible K / << M and M' K / . Thus, the dilatant
hardening effect vanishes in the limit that the pore fluid bulk modulus goes to
zero.
2.4 Discussion
Rice [1975] remarks that if initial material non-uniformities or variation of consti-
tutive parameters on the time scale of perturbation evolution are included, these
introduce additional inhomogeneous terms into (10). In the linearized analysis,
these terms are linear in the perturbation magnitudes and proportional to the
stress or strain rate of the uniform solution. For infinitesimal wavelengths, 0,
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 167

the growth rate in (12) is unbounded and these additional terms do not affect the
conditions for instability. In actuality, the perturbation wavelength is limited from
below by the material grain size lg . Consequently, the growth rate r is bounded
from above by Thus, Rices [1975] analysis corresponds to the
case when the perturbation growth rate is much larger than the layer deformation
rate, r max >> That is, perturbation growth occurs instantaneously on the time
scale of the layer deformation and the onset of perturbation growth ( c = 0) will
coincide with instability.
For very low permeability rocks (or fast loading rates), the diffusion length
scale l d = ( c / ) 1/2 may be comparable to the grain size, yielding rmax ~ In
this case, the actual instability, defined as a certain increase of the perturbation
magnitude over its initial value, will be delayed from the onset of perturbation
growth. To estimate the delay of undrained instability, Garagash and Rudnicki
[2000] have extended Rices analysis to include the coupling between the pertur-
bation and the evolution of the constitutive parameters with uniform background
deformation.
Rudnicki [1984b] examined the effect of an initial non-uniformity by considering
the shear of a weakened layer embedded in an infinite body. Both the layer and
the surrounding material deform non-elastically but the peak stress in the layer is
slightly less than that in the surrounding material. The pore pressure is assumed
to be uniform in both the layer and the surrounding material and the fluid mass
exchange is assumed to be proportional to the difference. The development of
instability in time depends on the ratio of the rate of imposed farfield strain-rate
to the rate of exchange of fluid mass between flow from the layer, d. In the
limit /d 0, the pore pressure in the layer is the same as in the surrounding
material and instability occurs at the peak of the drained stress-strain curve. For
finite /d, instability is delayed until the weakened layer reaches the peak of
its dilatantly hardened stress-strain curve. For small /d, as appropriate for
most applications, an asymptotic analysis predicts that the time delay is given
by where is the half-width of the peak of the stress-
strain curve, is the difference in the peak stresses of the weakened layer and
the surrounding material divided by times the elastic shear modulus and is a
nondimensional measure of the strength of dilatant hardening. These delay times
are less than a few hours for tectonic strain rates and less than a few tens of
seconds for typical laboratory strain-rates.
If dilatant hardening causes a sufficient reduction of the pore fluid pressure,
exsolution of dissolved gases or cavitation of the fluid may occur and the pore
fluid bulk modulus will be dramatically reduced. Equations (13) and (9) indicate
168 J. W. RUDNICKI

that the dilatant hardening effect diminishes with reduction of the pore fluid bulk
modulus and vanishes in the limit K 0. Rudnicki and Chen [1988] have
proposed that cavitation is a limit to strong dilatant hardening associated with
slip on a weakening frictional surface and suggest that this effect is consistent
with observations by Martin [1980] of pore fluid stabilization of rock failure. In
addition, in undrained biaxial experiments on a quartz sand, Mokni and Desrues
[1999] have observed that the formation of shear bands in dilatant specimens does
not occur until cavitation of the pore fluid. The numerical simulations of Schrefler
et al. [1996] also indicate the importance of cavitation in the formation of shear
bands.
3 Application to Compacting Materials
Although Rice [1975] considers only dilating materials ( > 0), he notes that
for loosely packed granular materials < 0. This will also be the case for high
porosity rocks that compact when sheared. For < 0, (8) indicates that the
effective compressive normal stress decreases and (9) that H u n d r a i n e d < H. T h i s
phenomenon could be described as compaction softening. Consequently, the de-
nominator of (11) will pass through zero before the numerator. The exponent
r in (12) reveals that spatial perturbations are damped infinitely fast and then
grow infinitely fast as H u n d r a i n e d , the denominator of (11), passes through zero.
This singular jump in diffusivity suggests that variation of constitutive parameters
with the uniform background deformation should be included in the perturbation
solution. By including this variation, Garagash and Rudnicki [2000] show that
undrained deformation of compacting material is stable (the perturbation magni-
tude vanishes) as H undrained passes through zero if the ratio of the diffusion length
l d = (c / )1/2 to the maximum perturbation wavelength (defined by the layer
width 2h) is larger than a critical value defined by the dependence of the inelastic
moduli H a n d H u n d r a i n e d on the deformation State:

(15)

This inequality indicates that the maximum perturbation wavelengths (h /2) are
most unstable for compaction softening in contrast to dilatant hardening for which
the shortest wavelengths grow most rapidly (12). This suggests that failure will
occur by a diffuse mode rather than a localized mode, as for dilatant hardening.
The stability criterion (15) implies an interesting scale effect: compaction soft-
ening is stable near the peak shear stress (H u n d r a i n e d = 0) if the specimen size h
is small enough. This result is consistent with the small scale laboratory results
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 169

of Han and Vardoulakis [1991] and Finno et al. [1995]. If the stability crite-
rion is not satisfied, the perturbation magnitude becomes algebraically singular at
H u n d r a i n e d = 0. Vardoulakis [1996b] has applied Rices analysis to study the sta-
bility of biaxial deformation of water-saturated sand and noted that the instability
at H undrained = 0 can also be mitigated by the introduction of rate-dependence in
the material constitutive behavior.
The discussion of compaction in the preceding paragraphs assumes > 0 and
the physical interpretation of as a friction coefficient would seem to preclude
the possibility that < 0. But, enters (1) as the local slope of the yield
surface. As sketched in Figure 2, the yield surface is open on the normal stress
axis. Consequently, purely normal stress does not cause inelastic deformation, as
appropriate for low porosity rocks. For loose soils or highly porous rock, inelastic
compaction (volume decrease) is due not only to inelastic shearing but may be
caused by purely hydrostatic (or in the one dimensional layer model here, by
purely normal) stress. In this case, the yield surface is closed on the normal stress
axis, as depicted in Figure 3. So-called cap models were introduced by Dimaggio
and Sandler [1971]. As sketched in Figure 3, < 0 is appropriate for this class of
models and the magnitude of becomes large as the axis is approached. Because
the stability condition contains only the product , the case of a material that
compacts under a stress state on the cap of the yield surface reduces to that
analyzed by Rice [1975].
The form of equation (3) assumes that all inelastic volume strain is associated
with inelastic shear strain. Consequently, no inelastic volume strain would occur
for loading by purely normal stress. For high porosity rocks and loose soils, this
will not be a complete description and a term contributing inelastic volume strain
for purely (effective) normal stress can be added to (3)

(16)
where k is an inelastic modulus that is the slope of a drained ( dp = 0) curve of
versus p at constant shear stress (d = 0). The effect of including this term
is to modify the elastic contribution to the volumetric strain in (4b) and in the
stability analysis to replace the elastic modulus M (or M ') by the effective value

(17)

For typical cases, k 0 but k << M " . Decreasing k will reduce the difference
between the drained and undrained hardening moduli. Thus, for small k, differ-
ences in drained versus undrained deformation may be difficult to observe unless
170 J. W. RUDNICKI

Figure 3: Schematic illustration of a cap on the yield surface. Note that


= d /d < 0.

the stress state is near enough to the axis so that values of are large and
negative.
Many geomaterials exhibit both compaction and dilation depending on the
initial confining stress, initial porosity and load path. Examples include simulated
fault gouge [Marone et al., 1990], loose sand [Finno et al., 1995] and limestone
[Baud et al., 2000]. When these materials are fluid-saturated and deformed without
allowing fluid flow from the boundaries, the evolution of localized zones depends
on the local rate of fluid flow, the imposed rate of straining, and the transition
from contraction to dilation, A simple analysis [Rudnicki, 1996] shows that small
variations in the evolution of porosity with shear strain can dramatically alter
the undrained response. Rudnicki et al. [1996] have suggested that compaction
softening followed by dilatant hardening may be an explanation for the evanescent
shear band structures observed in some experiments of Finno et al. [1995]. If
compaction softening causes the onset of localized deformation in a narrow zone
but gives way to dilatant hardening before full development of the band, formation
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 171

of a shear band in another orientation may occur when the dilatant hardening
response becomes unstable.
4 Generalization to Arbitrary Deformation States
In this section, I extend Rices [1975] analysis to arbitrary deformation states. The
analysis follows the lines of Rices [1976] general treatment of conditions for local-
ization of plastic deformation in rate-independent solids. That analysis is phrased
in terms of measures of stress and strain appropriate for arbitrary deformation
magnitudes and discusses localization both as a bifurcation from homogeneous
deformation and as the growth of an initial non-uniformity. Here, for simplicity, I
restrict attention to small strains and consider only small perturbations from com-
pletely uniform deformation. Runesson et al. [1998] have presented an analysis
of localization for undrained conditions for finite strains. Generalizations of the
present analysis to finite strains can be adapted from the treatments of Rudnicki
[1983], Runesson et al. [1998] or Coussy [1995].
4.1 Localization in Rate-independent Solids
Consider a homogeneous solid deforming in homogeneous fashion described by
a strain rate d 0 and stress-rate Perturbations from these uniform fields are
denoted by d and To investigate the possibility of localization in a planar
band the perturbed fields are taken to vary with distance from a plane with normal
n. The requirements of equilibrium and continuous velocities place the following
restrictions on the form of the perturbed fields [Rice, 1976]):

(18)
(19)
where g is a function of n x.
For nonlinear elastic or rate-independent elastic-plastic solids with a smooth
yield surface and plastic potential, the strain-rate can be given by an incrementally
linear relation:
(20)
where If K can be inverted, (20) can be written in the
alternative form:
(21)
where K and M are mutual inverses with components satisfying

(22)
172 J. W. RUDNICKI

and ij is the Kronecker delta. The components of the incremental compliances


K i j k l and the moduli M ijkl are symmetric with respect to interchange of the first
two indices and the last two indices, i.e., M i j k l = M j i k l and M i j k l = M i j l k , but
are not, in general, symmetric with respect to interchange of the first pair and the
last pair, i.e., M i j k l M klij . The last is typical of geomaterials for which inelastic
strain increments are not perpendicular to the yield surface or, in other words, the
plastic flow rule is not associated with the yield function. Substitution of (18) into
(21) and the result into (19) yields

(23)
where M 0 denotes the incremental moduli for the uniform fields. In the simplest
case, the constitutive response of the perturbed field is identical with that of the
uniform field and the right hand side of (23) vanishes. The first possibility for a
non-trivial solution for g occurs when the determinant of the coefficient matrix
vanishes:
(24)
w h e r e m jk = (n M n ) jk If the constitutive responses of the perturbed and
homogeneous fields differ, the condition (24) is still limiting in the sense that g
determined from (23) will be unbounded when (24) is met.
4.2 Inclusion of Pore Fluid
When the solid is saturated with a fluid that can be described in terms of a single
scalar pore pressure (see Cleary [1978] for a discussion of other possibilities), a
term involving the rate of pore pressure must be appended to (20):

(25)
Alternatively (25) can be written as

(26)
w h e r e = M : A. The introduction of the pore pressure as an additional field
variable requires an additional constitutive equation. This is conveniently taken
to be a relation the rate of change of fluid mass content per unit volume of porous
solid m = where is the density of homogeneous pore fluid and is apparent
void volume fraction. For arbitrarily compressible constituents, can be expressed
in terms of the rate of strain and pore pressure change:

(27)
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 173

or in terms of the stress-rate and pore pressure change by substituting (25) into
(27):

(28)
In (25), (26) and (27), A , and R are symmetric because of the symmetry of d
and
For drained response, the pore pressure is constant and (25) and (26) reduce
to (20) and (21). The condition for localization is given by (24).
For undrained response, the fluid mass in material elements is constant, the
right hand side of (27) is zero, and the pore pressure is given by

(29)
or, equivalently, in terms of the stress from (28)

(30)

Substitution of (29) into (26) gives

(31)

where the undrained incremental moduli are given by

(32)
The condition for localization (24) evaluated in terms of the undrained moduli
(32) is
(33)
where m u = n M u n. The matrix m u is related to m (24) by

(34)
where r = n R and a = n . The earlier discussion of the layer problem examined
by Rice [1975] suggests that (33) may be met before or after (24) depending on
the nature of the constitutive relation and, in particular, the inelastic volumetric
deformation. If (24) is met before (33), Rices [1975] analysis also suggests that
undrained homogeneous deformation is unstable, in the sense that small spatial
perturbations will grow exponentially in time. As a result, the condition (33) may
be irrelevant since instability may occur well before it is met.
174 J. W. RUDNICKI

4.3 Diffusive Instability


In order to analyze the stability of undrained homogeneous deformation, an equa-
tion expressing fluid mass conservation (in the absence of sources) must be added
to the field equations and an additional constitutive equation relating the fluid
mass flux to gradients in pore pressure must be specified. Fluid mass conservation
(in the absence of sources) is given by

(35)
where q is the mass flow rate per unit area of porous solid. Fluid mass flow is
assumed to be related to the pressure gradient by Darcys law:

(36)
where is the ratio of a symmetric permeability tensor to the (scalar) fluid viscos-
ity and 0 is the (constant) fluid mass density in the reference state. Substitution
of (36) into (35) yields
(37)
For simplicity, we assume, as in Rice [1975], that perturbations from undrained
homogeneous deformation are small enough so that incremental constitutive pa-
rameters are the same for the uniform and perturbed fields. If this is not the case,
additional inhomogeneous terms would enter the right hand sides of the equations.
As in (23), these terms involve the differences of the constitutive parameters mul-
tiplied by the strain-rate and pore pressure rate in the uniform field. Forming the
difference of (26) and substituting into (19) yields

(38)
where is the difference between the pore pressure-rates in the uniform and
perturbed fields and, again, a = n . Because the unperturbed fields are ho-
mogeneous and undrained, both sides of (37) are identically zero. The kinematic
restriction on the perturbed strain-rate field (18) suggests that p depends on
distance n x perpendicular to the orientation of a planar band and, hence, can
be decomposed into Fourier components of the form exp n x). This type of
deformation mode has also been discussed for sands by Vardoulakis [1996b]. For
this decomposition, using (27) in (37). Since both sides of (37) are zero for the
unperturbed (homogeneous, undrained) field, the difference fields also satisfy (37).
Using the Fourier decomposition on the right hand side and assuming the perme-
ability is uniform yields
(39)
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 175

where = n n. Forming the difference of (27) and substituting into (39)


yields
(40)
The vector g can be eliminated from (38) by writing the inverse of m as k / m
where k is the matrix of co-factors and m = det( m i j ). Substituting the result into
(40) yields
(41)
This is a linear ordinary differential equation for the time evolution of the per-
turbed pressure field. The solutions change from exponentially decaying to ex-
ponentially growing as m passes through zero from positive to negative values (at
least if the term in square brackets is positive). Since the vanishing of m is the
condition for localization in terms of the drained incremental moduli (24), this
condition controls the growth rate of spatial perturbations as in the layer analysis
of Rice [1975].
The similarity with Rices [1975] analysis suggests that the square bracket in
(41) will vanish when the condition for localization is met in terms of the undrained
moduli (33). Using m 1 = k /m in (34), pre- and post-multiplying by a and
dividing through by a = a a yields

(42)

Using this expression to eliminate m in (41) gives

(43)

If R = (so that a = r ) or if k is symmetric, then the second term in brackets


vanishes and the coefficient of is expressed in terms of the matrix entering
the localization condition in terms of the undrained moduli. Even in this case,
however, it is not clear that the sign of this term changes when (33) is met. For
geomaterials, M is generally not symmetric with respect to interchange of the
first and last pair of indices. Consequently, neither m nor k will be symmetric.
In addition, for the particular form of a constitutive relation discussed in the next
section, this lack of symmetry of M causes R and to differ.
5 Application to an Elastic-Plastic Relation
This section illustrates the analysis of the preceding section by specializing to a
particular form of elastic plastic constitutive relation. The strain-rate for plastic-
loading is given by
176 J. W. RUDNICKI

(44)
where the first term is the elastic contribution, C is the tensor of elastic compli-
ances, P is a tensor specifying the direction of inelastic strain increments in stress
space, Q is the tensor giving the normal to the yield surface in stress space, and h
is an inelastic hardening (softening) modulus. Deformation increments that tend
to make Q : 0 for h > 0 cause elastic unloading and for these the second
term is dropped. (If h < 0, increments tending to make Q : > 0 cause elastic
unloading). Thus, the tensor K in (20) is given by

(45)
The inverse of K, M in (21) is given by

(46)

where L = C 1 is the tensor of elastic moduli, p = L : P and q = L : Q .


The constitutive relation used by Rudnicki and Rice [1975] in their study of
localization has the form of (44) but the yield function and plastic potential depend
only on the first and second stress invariants. In this case, P and Q are given by

(47a)

(47b)
1/2
where N = s/2 , s = (1/3) tr I is the deviatoric stress tensor, = (s : s/2)
is the Mises equivalent shear stress, I is the identity tensor and and h a v e
interpretations similar to those in (1). If the elasticity is assumed to be isotropic
with shear and bulk moduli, G and K, respectively, and P and Q are given by
(50)

(48a)
(48b)
(48c)
Rudnicki [1984a, 1985] has extended the formulation to include pore fluid effects
in the form of (44) used by Rudnicki and Rice [1975]. He assumes isotropic
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 177

poroelasticity for the elastic strain increments and inelastic strain increments are
given by the second term of (44) with the stress replaced by the effective stress
+ p I . As already noted, much experimental evidence is consistent with this
form and Rice [1977] has argued that it is appropriate for inelasticity arising from
microcracking and frictional slip on surfaces with small real areas of contact. For
isotropic poroelasticity, Nur and Byerlee [1971] have shown that the proper form
of the effective stress is + p I. The Biot porous media parameter (often
denoted ) is equal to 1 where K is the elastic bulk modulus of the
porous solid and is an additional modulus related to the bulk modulus of the
solid constituents [Rice and Cleary, 1976]. Thus, if the solid constituents are
incompressible, and = 1 This is a suitable approximation for soils, in
which the compressibility of the grains is much less than that of the porous matrix.
Under the same conditions for which + p I is the form of the effective stress for
inelastic deformation, Rice [1977] has also argued that the inelastic increment of
the apparent void volume fraction where = m / ) is equal to the inelastic
volume strain increment. Under these circumstances, the tensors A in (25) and
in (26) are given by

(49a)

(49b)
The tensor R and the coefficient entering (27) are given by

(50a)

(50b)
where B is Skemptons coefficient, the negative of the ratio of pore pressure to
mean normal stress during undrained elastic deformation [Rice and Cleary, 1976].
The rate of change of pore pressure during undrained deformation (30) is given
by
(51)
For isotropic elasticity, C : I = I : C =(1/3 K)I and Q : L : P = G +tr(Q ) tr(P)K .
The compliance tensor for undrained deformation can be formed from these
results

(52)
178 J. W. RUDNICKI

where hu n d = h + BK tr P tr Q / and the primes on P and Q denote the deviatoric


parts. For isotropic elasticity, the first two terms correspond to replacing K ,
the elastic bulk modulus for drained deformation by K u the value for undrained
deformation, where
(53)
The contribution to the inelastic strain-rate, the last term in (52), is identical in
form to that for drained deformation (44), with h replaced by h und , tr P replaced
by (1 B ) tr P and tr Q by (1 B ) tr Q. If the expression for the compliance is
specialized to pure shear, the result is exactly analogous to that of Rice [1975]:
The undrained response is identical in form to that for drained but with h replaced
by h und . Since B, K, and 1 are positive, h und is greater or less than h depending
on the sign of tr P tr Q, in a manner identical to the dependence on the sign of
in Rice [1975].
These expressions can also be used to examine the role of the compressibilities
of the individual solid and fluid constituents. If the solid constituents are incom-
pressible, = 1, and = R = I. Consequently, the results simplify considerably.
In particular a = r in equation (34) and the term r k a a k r vanishes in
(43).
The compressibility of the pore fluid enters only through the Skempton coeffi-
cient B . For isotropic poroelasticity, B can be expressed as

(54)

where 0 is the apparent void volume fraction in the reference state, K is the
pore fluid bulk modulus, and is another modulus related to the bulk modulus
of the solid constituents [Rice and Cleary, 1976]. If the solid constituents are
incompressible, If the fluid is also incompressible,
t h e n B = 1. In this limit, the bulk modulus is eliminated from (52) and so is
tr P. Consequently, both the elastic and inelastic components are incompressible,
consistent with the limit of incompressible constituents. Furthermore, tr Q is elim-
inated from (52). Therefore, any dependence of the yield function on mean stress
for drained response is exactly compensated by changes in pore pressure during
undrained response. Thus, the undrained response is incompressible and does not
depend on the mean stress. For such a solid, shear bands are predicted to occur at
45 from the principal axes, at least if P' = Q ' . Runesson et al. [1996] obtain this
result by direct calculation for undrained deformation of a porous elastic plastic
solid with an incompressible pore fluid. The result pertains, however, only when
the solid constituents are also incompressible. Although Runesson et al. [1996]
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 179

do not explicitly assume that the solid constituents are incompressible, they use
+ p I as the effective stress for both inelastic and elastic deformation. As dis-
cussed earlier, this form is consistent with linear poroelasticity only when the solid
constituents are incompressible.
6 Concluding Discussion
Application of Rices [1975] results to materials that compact with inelastic shear-
ing suggest that the accompanying pore pressure changes will be destabilizing.
That is, the effective hardening modulus for undrained response will be less than
that for drained response and conditions for localization will be met for undrained
constitutive response before they are met in terms of the drained response. But
this conclusion appears to apply only when the yield stress in pure shear increases
with increasing mean stress. For highly porous materials that compact with mean
stress, the yield surface typically has a cap and, as a result, the yield stress in pure
shear decreases with increasing compressive mean stress. In this case, the condi-
tions for linearized stability of undrained deformation revert to those obtained by
Rice [1975]: undrained deformation is stiffer than the drained response, but ho-
mogeneous undrained response becomes unstable when conditions for localization
are met in terms of the drained response.
The conclusions of Rice [1975] based on analysis of a layer subject to simple
shear and compression are shown to apply to arbitrary deformations, although
conditions for localization are not, in general, met at the peak of the stress versus
strain curve as they are for the layer. Spatial perturbations from the undrained so-
lution grow exponentially when the condition for localization is met in terms of the
drained response. In the layer problem, the growth rate is unbounded immediately
after the condition for localization is met in terms of the undrained response. This
is not necessarily the case for arbitrary three dimensional deformations because of
the lack of symmetry in the constitutive tensors that is typical of geomaterials.
For elastic plastic solids, the correspondence with the simple layer analysis is
direct. The effective hardening modulus governing shear for undrained conditions
is the sum of the value for drained response and a term that is the product of an
elastic bulk modulus and the mean parts of tensors giving the plastic flow direction
and the normal to the yield surface. The incremental compliance for undrained
deformation of an elastic plastic solid gives some insight into the surprising result
of Runesson et al. [1996]: if the pore fluid is completely incompressible, zones of
localization are predicted to occur at 45 to the principal axes regardless of the
form of the yield function and plastic potential (at least, if the deviatoric portions
of the tensors giving the plastic flow direction and the normal to yield surface
180 J. W. RUDNICKI

coincide). This results because, if the solid constituents are also incompressible,
undrained response is completely incompressible and the induced changes in pore
pressure exactly compensate for any dependence of the yield function on the mean
stress.
Acknowledgement. I am grateful to Dmitry Garagash for many helpful
discussions. I also wish to acknowledge interesting correspondence of several
years ago with Stein Sture and Kenneth Runesson on conditions for localization
for undrained deformation of fluid-saturated porous solids. Partial financial
support for this work was provided by the U. S. Dept. of Energy, Office of
Basic Energy Sciences, Geosciences Research Program through Grant No.
DE-FG02-93ER14344/09 to Northwestern University.

References
Baud, P. Alexandre Schubnel and T. -F. Wang, Dilatancy, compaction and failure model
in Solnhofen limestone, J. Geophys. Res., in press, 2000.
Brace, W. F. and R. J. Martin, III, A test of the law of effective stress for crystalline
rocks of low porosity, Int. J. Rock Mech. Mining Sci., Vol. 5, 415-436, 1968.
Cleary, M. P. Elastic and dynamic response regimes of fluid-impregnated solids with
diverse microstructures, Int. J. Solids Structures, 795-819, 1978.
Coussy, O. Mechanics of Porous Media, John Wiley and Sons, Ltd., Chichester, 1995.
Dimaggio, F. L. and I. S. Sandler, Material model for granular soils, J. Eng. Mech. Div.
ASCE, Vol. 97, 935-950, 1971.
Finno, R. J., W. W. Harris, M. A. Mooney, and G. Viggiani, Shear bands in plane strain
compression of loose sand, Gotechnique, Vol. 47, 149-165, 1997.
Garagash, D. and J. W. Rudnicki, Stability of undrained deformation of fluid-saturated
dilating/compacting solids (Abstract), 20th Int. Cong. of Theor. and Appl. Mech.,
Chicago, II, Aug. 27 - Sept. 2, 2000.
Han, C. and I. G. Vardoulakis, Plane strain compression experiments on water-saturated
fine-grained sand, Gotechnique, Vol. 41, 49-78, 1991.
Issen, K. A. and J. W. Rudnicki, Conditions for compaction bands in porous rock, J.
Geophys. Res., in press, 2000.
Marone, C., C. B. Raleigh, and C. H. Scholz, Frictional behavior and constitutive mod-
eling of simulated fault gouge, J. Geophys. Res., Vol. 95, 7007-7025, 1990.
Martin, R. J. III, Pore pressure stabilization of failure in Westerly granite, Geophys. Res.
Letters, Vol. 7, 404-406, 1980.
Mokni, M. and J. Desrues, Shear localization measurements in undrained plane-strain
biaxial tests on Hostun RF sand, Mech. of Cohesive-Frictional Materials, Vol. 4,
DIFFUSIVE INSTABILITIES IN GEOMATERIALS 181

419-441, 1999.
Mollema, P. N. and M. A. Antonellini, Compaction bands: a structural analog for anti-
mode I cracks in aeolian sandstone, Tectonophysics, Vol. 267, 209-228, 1996.
Nur, A. and J. D. Byerlee, An exact effective stress law for elastic deformation of rock
with fluids. J. Geophys. Res., Vol. 76, 6414-6419, 1971.
Olsson, W. A., Theoretical and experimental investigation of compaction bands, J. Geo-
phys. Res., Vol. 104, 7219-7228, 1999.
Paterson, M. S. Experimental Rock Deformation: The Brittle Field, New York: Springer-
Verlag, 1978.
Reynolds, O., On the dilatancy of media composed of rigid particles in contact, with ex-
perimental illustrations, Phil. Mag. (reprinted in Papers on Mechanical and Physical
Subjects by O. Reynolds, Cambridge University Press, New, York, 1901, Vol. 2, pp.
203-216), 1885.
Rice, J. R. On the stability of dilatant hardening for saturated rock masses. J. Geophys.
Res., Vol. 80, 1531-1536, 1975.
Rice, J. R. The localization of plastic deformation, in Proceedings of the 14th Int. Union
Theor. and Applied Mech. Congress, ed. W. T. Koiter, pp. 207-220, North Holland,
Amsterdam, 1976.
Rice, J. R. Pore pressure effects in inelastic constitutive formulations for fissured rock
masses. In Advances in Civil Engineering through Engineering Mechanics, pp. 360-
363. New York: American Society of Civil Engineers, 1977.
Rice, J. R. and M. P. Cleary, Some basic stress diffusion solutions for fluid-saturated
elastic porous media with compressible constituents, Rev. Geophys. Space Phys., Vol.
14, pp. 227-241, 1976.
Rudnicki, J. W. and J. R. Rice, Conditions for the localization of deformation in pressure-
sensitive dilatant materials. J. Mech. Phys. Solids, Vol. 23, 371-394, 1975.
Rudnicki, J. W. A formulation for studying coupled deformation - pore fluid diffusion
effects on localization. In Geomechanics, Proceedings of the Symposium on the Me-
chanics of Rocks, Soils and Ice, Applied Mechanics Division, Vol. 57 (edited by S.
Nemat-Nasser), pp. 35-44, American Society of Mechanics Engineers, New York, 1983.
Rudnicki, J. W. A class of elastic-plastic constitutive laws for brittle rock, J. of Rheology,
Vol. 28, 759-778, 1984a.
Rudnicki, J. W. Effects of dilatant hardening on the development of concentrated shear
deformation in fissured rock masses, J. Geophys. Res., Vol. 89, 9259-9270, 1984b.
Rudnicki, J. W. Effect of pore fluid diffusion on deformation and failure of rock, in
Mechanics of Geomaterials (edited by Z. P. Ba ant), pp. 315-347, John Wiley &
Sons, Ltd., New York, 1985.
Rudnicki, J. W. and C.-H. Chen, Stabilization of rapid frictional slip on a weakening
182 J. W. RUDNICKI

fault by dilatant hardening, J. Geophys. Res., Vol. 93, 4745-4757, 1988.


Rudnicki, J. W., R. J. Finno, M. A. Alarcon G. Viggiani, and M. A. Mooney, Coupled
deformation-pore fluid diffusion effects on the development of localized deformation in
fault gouge, in Predictions and Perfomance in Rock Mechanics and Rock Engineering,
EUROCK96, edited by G. Barla, pp.1261-1268, Balkema, 1996.
Runesson, K., D. Peri and S. Sture, Effect of pore fluid compressibility on localization
in elastic-plastic porous solids under undrained conditions, Int. J. Solids Structures,
Vol. 33, 1501-1518, 1996.
Runesson, K. R. Larsson, and S. Sture, Localization in hyperelasto-plastic porous solids
subjected to undrained conditions, Int. J. Solids Structures, Vol. 35, 4239-4255, 1998.
Schrefler, B. A., L. Sanavia and C. E. Majorana, A multiphase medium model for locali-
sation and post localization simulation in geomaterials, Mech. of Cohesive-Frictional
Materials, Vol. 1, 95-114, 1996.
Vardoulakis, I. Stability and bifurcation of undrained, plane rectilinear deformations on
water-saturated granular soils, Int. J. Num. and Anal. Meth. Geomech., Vol. 9, 399-
414, 1985.
Vardoulakis, I. Dynamic stability analysis of undrained simple shear on water-saturated
granular soils, Int. J. Num. and Anal. Meth. Geomech., Vol. 10, 177-190, 1986.
Vardoulakis, I. Deformation of water-saturated sand: I. uniform undrained deformation
and shear banding, Gotechnique, Vol. 46, 441-456, 1996a.
Vardoulakis, I. Deformation of water-saturated sand: II. effect of pore water flow and
shear banding, Gotechnique, Vol. 46, 457-472, 1996b.
Wang, T.-F., C. David, and W. Zhu, The transition from brittle faulting to cataclastic
flow in porous sandstones: Mechanical deformation, J. Geophys. Res., Vol. 102, 3009-
3025, 1997.
FRACTURE MECHANICS OF AN INTERFACE CRACK
BETWEEN A SPECIAL PAIR OF TRANSVERSELY
ISOTROPIC MATERIALS

LESLIE BANKS-SILLS AND VINODKUMAR BONIFACE


The Dreszer Fracture Mechanics Laboratory
Department of Solid Mechanics, Materials and Structures
The Fleischman Faculty of Engineering
Tel Aviv University
69978 Ramat Aviv, Israel

Abstract. In this investigation, the Stroh formulation is employed to develop the stress
and displacement fields in the vicinity of an interface crack between two specially oriented
transversely isotropic materials. The lower material is mathematically degenerate. In ad-
dition, a conservative integral is employed in conjunction with the finite element method
to calculate stress intensity factors. The derived stress and displacement fields are used
as auxiliary fields in the M-integral for extraction of the stress intensity factors. As a
benchmark problem for this calculation, the asymptotic displacements are prescribed on
the boundary of a circular domain. Excellent numerical results are obtained.

1. Introduction

The mechanics of interface cracks between two joined materials is an im-


portant subject which has received much attention in the literature. The
papers by J.R. Rice (1988) and J.W. Hutchinson (1990) presented a new
view on the subject which delineated a philosophy for tackling the problem
of a crack between two isotropic materials. One of the controversial issues
which hindered progress was treatment of crack face contact which naturally
spreads from the crack tip along the crack faces. The new approach was to
neglect this region when it is sufficiently small and within the small scale
yielding zone. In this situation, approximate bounds for the phase angle
(mode mixity) were determined by Rice (1988). It was found that there
was a large range of for which the contact region is sufficiently small.
With this approach, interest in this subject was rekindled and a wealth of
investigations followed.
183
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 183204.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
184 L. BANKS-SILLS AND V. BONIFACE

Following in this direction, Suo (1990) considered an interface crack


between two dissimilar anisotropic materials for both non-oscillatory and
oscillatory singularities. Investigations were carried out earlier for the case
of a non-oscillatory singularity. In particular, Bassani and Qu (1989) and Qu
and Bassani (1989) derived a condition for non-oscillatory behavior. In Suo
(1990), the structure of the near tip field was determined for the oscillatory
singularity; it is similar to that of a crack between two isotropic materials.
Working independently, Ting (1986, 1990) determined the singularity for
both the oscillatory and non-oscillatory cases. In addition, he developed
the asymptotic stress and displacement fields (Ting, 1992).
In order to employ a fracture mechanics approach for predicting the
propagation of cracks between two anisotropic materials, stress intensity
factors and a fracture criterion are required. In this investigation, the Stroh
formulation is employed to derive the stress and displacement fields in the
neighborhood of the tip of an interface crack between two transversely
isotropic materials. An excellent treatise on these issues is provided by
Ting (1996). The upper material is taken so that its symmetric plane is
perpendicular to the x 1 direction, whereas the symmetric plane of the
lower material is perpendicular to the x 3 direction. Of course, within each
of these planes, there is no preferential direction. The symmetric plane of
the lower material causes a mathematical degeneracy requiring an analysis
which is more complicated than usual.
In addition, a method for calculating stress intensity factors from finite
element results is employed. This method has already been described and
employed for isotropic bimaterials (see Wang and Yau, 1981, Shih and
Asaro, 1988, Matos, et al., 1989, Nakamura, 1991, Nahta and Moran, 1993,
Gosz, et al., 1998, and Banks-Sills, et al., 1999) and certain orthotropic
bimaterials (Charalambides and Zhang, 1996). The M-integral, developed
by Wang and Yau (1981), allows for the determination and separation of
the stress intensity factors K 1 and K 2 for interface cracks between two
isotropic materials. Because of the similar structure of the near tip fields
for cracks between two anisotropic materials, the same integral formulation
may be employed. This was done by Charalambides and Zhang (1996) for
two general orthotropic materials. There are cases however, for which the
near tip stress and displacement fields in that study may not be employed.
For certain material symmetries, there is a mathematical degeneracy which
invalidates these fields and requires a more complicated analysis. In this
study, a particular material direction is taken which leads to this degenerate
situation.
In Section 2, the stress and displacement fields will be described for
the particular materials chosen for study. The M-integral is presented in
Section 3. A benchmark problem is carried out in Section 4 to demonstrate
FRACTURE MECHANICS OF AN INTERFACE CRACK 185

accuracy of the method.

2. Stress and Displacement Fields

Relevant concepts related to interface fracture for both isotropic and aniso-
tropic materials are presented here. In two dimensions and referring to
Fig. 1, the in-plane stresses in the neighborhood of a crack tip at an interface
are given by

(1)

where , = 1, 2, i = the complex stress intensity factor

K = K 1 + i K2 , (2)
and the superscripts (1) and (2) are related to the real and imaginary parts

Figure 1. Crack tip coordinates.

of Kr i , respectively. In (1), following Ting (1996)

(3)

where
(4)

The 3 3 matrix is given by

(5)
(6)
and
(7)
186 L. BANKS-SILLS AND V. BONIFACE

The subscripts 1 and 2 in (6) and (7) represent, respectively, the upper and
lower material. Since Sj and L j are real and

(8)
knowledge of the left hand side of (8) is sufficient to determine (6) and
(7). Instead of presenting more general expressions for the matrices Aj and
B j which may be found in Ting (1996, pp. 170172), the specific matrices
for the upper and lower materials are presented in Appendix 1. Note that
transversely isotropic materials are being studied here. It is observed that
the notation employed here for in (3) differs from that of Rice (1988) and
Hutchinson (1990) in which the fraction is inverted. The notation here is in
keeping with Dundurs (1969). However, it is immaterial if one is consistent.
Referring again to the expression for the stresses in (1), the functions
and are presented in Appendix 2 for the two types of materials
considered in this study. For two isotropic materials they are given in polar
coordinates by Rice, et al. (1990) and in Cartesian coordinates by Deng
(1993).
The complex stress intensity factor in (2) may be written in non-dimen-
sional form as

(9)
where L is an arbitrary length parameter and is the applied stress. The
non-dimensional complex stress intensity factor may be written as

(10)
so that the phase angle

(11)

In (11)

(12)

(13)

The material parameters D 11 and D 22 always have the same sign; they
are components of the matrix D in (6). The mechanical properties E A ,
E T , G A , G T , v A and v T are the usual material properties in the axial and
FRACTURE MECHANICS OF AN INTERFACE CRACK 187

transverse directions (namely, Youngs moduli, shear moduli and Poissons


ratios); since the material is transversely isotropic, G T = E T /2(1+v T ). The
parameter is given in Appendix 1 in eq. (44). The constants j , j = 1, 2, 3
are related to the three complex eigenvalues of the elastic constants p j for
the upper material (see Ting, 1996, pp 121128), where p j = i j for a
transversely isotropic material with this material symmetry.
For plane strain conditions, the stress components on the interface ahead
of the crack tip are

(14)

The crack face displacements in the vicinity of the crack tip are found to
be

(15)

where
The interface energy release rate G i is related to the stress intensity
factors by

(16)

where

(17)

Note that the subscript i in (16) represents interface and G i has units of
force per length.
It should be noted that inherently for any interface both K 1 and K 2
must be prescribed or equivalently G i and . In describing an interface
crack propagation criterion, one may prescribe a relation between K 1 and
K 2 or what is commonly done, the critical energy release rate G i c is given
as a function of the phase angle .
A possible generalization of the present problem is one in which the
crack in Fig. 1 is rotated about the x 3 -axis while ensuring that the sym-
metry plane of the upper material remains perpendicular to the crack line.
For this case, it may be observed from (4) that remains invariant as was
pointed out earlier by Ting (1986). Moreover, the complex stress intensity
factor K can be defined as in (14) based upon a new crack tip coordinate
system aligned along and normal to the crack line. The resulting stress and
188 L. BANKS-SILLS AND V. BONIFACE

displacement fields can be transformed to the original x 1 -x 2 coordinate


system using standard transformation matrices.

3. Area M-Integral

In this section, the path independent M-integral is described. The M-


integral was introduced by Yau, et al. (1980) for separation of mixed modes
in homogeneous bodies. It was converted to an area integral for isotropic
bimaterials by Shih and Asaro (1988). It may be written as

(18)

In (18), indicial notation is employed, the superscripts (1) and (2) represent
two solutions and is the Kronecker delta. The mutual strain energy density
W (1,2) of the two solutions is given by

(19)
The function q 1 is defined for finite element analysis as

(20)

where N m are the finite element shape functions of an eight noded isopara-
metric element and and are the coordinates in the parent element (for
further details, see Banks-Sills and Sherman, 1992). The calculation of the
M-integral is carried out in a ring of elements surrounding the crack tip
(the area A in (18)). The elements within the ring move as a rigid body. For
each of these elements q 1 is unity; so that, the derivative of q 1 with respect
to x j is zero. For all elements outside the ring, q 1 is zero; so that, again
the derivative of q 1 is zero. For elements belonging to the ring, the vector
q 1m in (20) is chosen so that the virtual crack extension does not disturb
the relative nodal point positions in their new locations; for example, a
regular element with nodes at the mid-sides contains only mid-side nodes
after distortion. The relationship

(21)
may also be obtained.
In (18) and (21), problem (1) is that for which a solution is sought. Two
auxiliary solutions are required in order to determine both and
for this problem. These are denoted as (2a) and (2b).
FRACTURE MECHANICS OF AN INTERFACE CRACK 189

For solution (2a), choose = 1 and = 0. Such a solution does


exist for some special loading. Equation (21) becomes

(22)
and from (18)

(23)

The displacements required for solution (1) are taken from a finite element
analysis of the problem to be solved; the stresses and strains are calculated
from these. Asymptotic expressions for the stresses, strains and displace-
ments for solution (2a) are employed. The stresses and displacements are
presented in Appendix 2.
For solution (2b ) , = 0 and = 1. Equation (21) becomes

(24)

and from (18)

(25)

After calculating these integrals, and are found by equating


(22) and (23), (24) and (25).

4. Benchmark Problem

A benchmark problem is presented in this section. Here, a circular domain


containing a crack along the interface between two transversely isotropic
materials is analyzed under plane strain conditions (see Fig. 2). The mate-
rial chosen here is a graphite/epoxy (AS4/3501-6) fiber reinforced material.
The volume fraction of the fibers is about 65%. Although the graphite fiber
is anisotropic, homogenization allows determination of the effective mechan-
ical properties, some of which are presented in Table 1 (Pagano, 1999). In
the upper material, the graphite fibers are aligned with the x1 direction,
while in the lower material they are along the x3direction resulting in a
mathematical degeneracy. Parameters necessary for calculation of the stress
intensity factors and energy release rate are given in Table 2.
190 L. BANKS-SILLS AND V. BONIFACE

TABLE I. Effective me-


chanical properties of a
graphite/epoxy fiber re-
inforced composite.

property value
EA 138.2 GPa
ET 10.4 GPa
GA 5.5 GPa
vA 0.3
vT 0.55

TABLE II. Material


parameters required for
stress intensity factor
and energy release rate
calculation.

parameter value
1 0.624
2 4.896
3 1.280
-0.028
D11 (GPa) 1 0.231
D 22 (GPa) 1 0.312

For the geometry in Fig. 2, two analyses are performed. In the first case,
K 1 is taken to be unity and K 2 is taken as zero. The asymptotic displace-
ment field corresponding to these values (calculated from the expressions
presented in Appendix 2) is applied to the outer boundary of the circular
domain. Finite element analyses are carried out with the program ADINA
(1999). The finite element mesh is exhibited in Fig. 3. There are 4,800
eight noded isoparametric elements and 14,641 nodal points. Quarter-point
elements are employed at the crack tip. Although the singularity at the
crack tip is a combination of square root and oscillatory, it was found by
Banks-Sills, et al. (1999) that better results are obtained for interface cracks
in bimaterial isotropic bodies when quarter-point elements are employed
instead of regular eight noded isotropic elements. The result for K1 is found
to be 1.0 + K 1 where 2 10 5 K 1 3 10 4 for eight rings not
FRACTURE MECHANICS OF AN INTERFACE CRACK 191

Figure 2. Crack in a transversely isotropic, bimaterial circular domain.

including the innermost two. For K 2 , the value was found to be between
2 1 0 5 and 6 1 0 5 .
As the second case, the displacement field corresponding to K 1 = 0 and
K 2 = 1 is applied at the boundary. The results obtained by means of the
M-integral are 8 10 5 K 1 5 10 5 and K 2 = 1.0 + K 2 w h e r e
3 10 4 K 2 5 10 4. In this case, there is crack face overlap.
The results of these two load cases demonstrate the accuracy of the
stress and displacement fields presented in Appendix 2, as well as the M-
integral implementation.

5 . Summary and Discussion

The first term of an asymptotic expansion for both the stress and dis-
placement fields has been developed for a crack along the interface of
two transversely isotropic materials. The plane of symmetry for the upper
material is perpendicular to the x 1 direction, whereas that for the lower
material is perpendicular to the x 3 direction. Thus, the lower material
involves a mathematical degeneracy.
These fields have been employed as auxiliary functions in a conservative
area M-integral. A benchmark problem was solved and shown to produce
excellent results.
Work is continuing in this direction for other cases with transverse
isotropy. Some analytic solutions will be reported elsewhere.

Acknowledgment

The first author would like to give special thanks to Professor Anthony R.
Ingraffea who provided a wonderful environment at Cornell University to
carry out a large portion of this investigation while she was on sabbatical
192 L. BANKS-SILLS AND V. BONIFACE

Figure 3. Finite element mesh of circular domain containing 14,641 nodal points and
4,800 eight noded isoparametric elements.

there. We would also like to thank Mr. Rami Eliasi for all his help with the
finite element calculations.

References

Banks-Sills, L. and Sherman, D.(1992) On the computation of stress intensity factors


for three-dimensional geometries by means of the stiffness derivative and J-integral
methods, International Journal of Fracture 53, 120.
Banks-Sills, L., Travitzky, N., Ashkenazi, D. and Eliasi, R. (1999) A methodology for
measuring interface fracture properties of composite materials, International Journal
of Fracture 99, 143161.
Bassani, J. and Qu, J. (1989) Finite crack on bimaterial and bicrystal interfaces. Journal
of the Mechanics and Physics of Solids 37, 434453.
Bathe, K.J., (1999) ADINA Automatic Dynamic Incremental Nonlinear Analysis
System, Version 7.3, Adina Engineering, Inc. USA.
FRACTURE MECHANICS OF AN INTERFACE CRACK 193

Charalambides P.G. and Zhang, W. (1996) An energy method for calculating the stress
intensities in orthotropic bimaterial fracture, International Journal of Fracture 76,
97120.
Deng, X. (1993) General crack-tip fields for stationary and steadily growing interface
cracks in anisotropic bimaterials. Journal of Applied Mechanics 60, 183189.
Dundurs, J. (1969) Edge-bonded dissimilar orthogonal elastic wedges under normal and
shear loading, Journal of Applied Mechanics 36, 650652.
Gosz, M., Dolbow, J. and Moran, B. (1998) Domain integral formulation for stress inten-
sity factor computation along curved three-dimensional interface cracks, International
Journal of Solids and Structures 35, 17631783.
Hutchinson, J.W. (1990) Mixed-mode fracture mechanics of interfaces, in M. Rhle,
A.G. Evans, M.F. Ashby, J.P. Hirth (eds.), Metal-Ceramic Interfaces, Pergamon Press,
Oxford, 295301.
Matos, P.P.L., McMeeking, R.M., Charalambides, P.G. and Drory, M.D. (1989) A method
for calculating stress intensities in bimaterial fracture, International Journal of Fracture
40, 235254.
Nahta, R. and Moran, B. (1993) Domain integrals for axisymmetric interface crack
problems, International Journal of Solids and Structures 30, 20272040.
Nakamura, T. (1991) Three-dimensional stress fields of elastic interface cracks, Journal
of Applied Mechanics 58, 939946.
Pagano, N. (1999) Personal communication.
Qu, J and Bassani J.L. (1989) Cracks on bimaterial and bicrystal interfaces. Journal of
the Mechanics and Physics of Solids 37, 417434.
Rice, J.R. (1988) Elastic fracture mechanics concepts for interfacial cracks, Journal of
Applied Mechanics 55, 98103.
Rice, J.R., Suo, Z., and Wang, J.-S. (1990) Mechanics and thermodynamics of brittle
interface failure in bimaterial systems. in M. Rhle, A.G. Evans, M.F. Ashby, J.P.
Hirth (eds), Metal-Ceramic Interfaces, Pergamon Press, Oxford, 269294.
Shih, C.F. and Asaro, R.J. (1988) Elastic-plastic analysis of cracks on bimaterial
interfaces: part Ismall scale yielding, Journal of Applied Mechanics 55, 299316.
Stroh, A.N. (1958) Dislocations and cracks in anisotropic elasticity, Philosophical
Magazine 7, 625646.
Suo, Z. Singularities, interfaces and cracks in dissimilar anisotropic media, Proceedings
of the Royal Society, London 427, 331358.
Ting, T.C.T. (1986) Explicit solution and invariance of the singularities at an interface
crack in anisotropic composites, International Journal of Solids and Structures 22,
965983.
Ting, T.C.T. and Hwu, C. (1988) Sextic formalism in anisotropic elasticity for almost
non-semisimple matrix N, International Journal of Solids and Structures 24, 6576.
Ting, T.C.T. (1990) Interface cracks in anisotropic bimaterials, Journal of the Mechanics
and Physics of Solids 38, 505513.
Ting, T.C.T. (1992) Interface cracks on anisotropic elastic bimaterialsa decomposition
principle, International Journal of Solids and Structures 29, 19892003.
Ting, T.C.T. (1996) Anisotropic ElasticityTheory and Applications, Oxford University
Press, Oxford.
Wang, S.S. and Yau, J.F. (1981) Interfacial cracks in adhesively bonded scarf joints,
American Institute of Aeronautics and Astronautics Journal 19, 13501356.
Yau, J.F., Wang, S.S. and Corten, H.T. (1980) A mixed-mode crack analysis of isotropic
solids using conservation laws of elasticity. Journal of Applied Mechanics 47, 335341.
194 L. BANKS-SILLS AND V. BONIFACE

Appendix 1

In this Appendix, the matrices A j , B j and B j 1 which appear in (8) and


are employed to calculate the oscillating part of the singularity in (3)
are presented for the specific transversely isotropic materials studied here.
They are related to the material properties. The subscript j represents the
upper and lower materials, 1 and 2, respectively.
The x 1 direction is the axial direction of the upper material. The
matrix A 1 is given by

(26)

where , j = 1,2,3, are normalization factors for the upper material


which are not necessary in the calculation of (8). The constants j , j =
1,2,3 are related to the three complex eigenvalues of the elastic constants
p j (see Ting, 1996, pp 121128), where p j = i j for a transversely isotropic
material with this material symmetry. They are given by

(27)

(28)

where are elements of the reduced compliance matrix, which for the
present material are found to be

(29)

(30)

(31)

(32)

(33)
FRACTURE MECHANICS OF AN INTERFACE CRACK 195

The constants Q j are related to the material properties as

(34)

(35)

(36)

(37)

The material parameters E A , E T , G A , GT , vA and v T are the usual material


properties in the axial and transverse directions (namely, Youngs mod-
uli, shear moduli and Poissons ratios); since the material is transversely
isotropic, G T = E T / 2 ( 1 + v T ).
The matrix B 1 is given by

(38)

Its inverse is given by

(39)

In the lower material, the axial direction coincides with the x 3 direction.
The mechanical properties E A , E T , G A , G T , v A and v T are taken to be the
same as for the upper material; but they are in different coordinate direc-
tions. It turns out that this material is mathematically degenerate. It has
three identical complex eigenvalues p j = i where the subscript j = 1,2,3.
To determine the stress and displacement fields, matrices alternative to A 2
and B 2 are required; these are A '2 and B ' 2 . Since

(40)
it is possible to calculate with the aid of (8). On the other hand, one may
determine A 2 B 2 1 without calculating the individual matrices (see Ting,
1996, p. 173).
For brevity, only the primed matrices are presented. To obtain them,
an orthogonalization procedure is employed; for details see Ting (1996,
pp. 489492) and Ting and Hwu (1988). They are found to be
196 L. BANKS-SILLS AND V. BONIFACE

(41)

(42)

and

(43)

where

(44)

The normalization factors and are again unnecessary for deter-


mining both and the stress and displacement fields. Nonetheless, for this
special case of a mathematically degenerate material, they are seen to be

(45)

(46)

Appendix 2

In this Appendix, the stress and displacement fields in the neighborhood


of the crack tip are derived for the particular material combination con-
sidered in this investigation. Recall that both materials are transversely
isotropic. In the upper material, the axial direction coincides with the
x 1 direction; whereas for the lower material, this direction coincides with
the x 3 direction. However, for any anisotropic material in the upper half-
plane and mathematically degenerate material in the lower half-plane, the
displacement field u ( j ) and the stress function ( j ) are presented below
( j = 1,2 represents upper and lower material, respectively). Both the
FRACTURE MECHANICS OF AN INTERFACE CRACK 197

displacement and stress function are two-dimensional vectors in the x1 a n d


x 2 directions.
For the upper half-plane,

(47)

(48)

where the diagonal matrix is

(49)

and h is a complex 3 1 vector to be determined.


Equations (47) and (48) are a special case of equations found in Ting (1992).
For the specific case at hand, transversely isotropic material with the axial
direction coinciding with the x1 axis, p j = i j , j = 1, 2, 3. The bar over a
quantity represents its complex conjugate. The matrices A 1 and B 11 a r e
given in Appendix 1.
For a mathematically degenerate anisotropic material in the lower half-
plane,

(50)

(51)

where the matrix F ( z ) is given by

(52)

and ()' represents differentiation with respect to z. For a


transversely isotropic material with the axial direction coinciding with the
x 3 axis, p j = i ; hence, z ( 2 ) = z = x 1 + ix 2 . The vector h is determined by
satisfying displacement continuity across the interface and is found to be

(53)

where

(54)
198 L. BANKS-SILLS AND V. BONIFACE

From the above relations, explicit expressions for the stress and dis-
placement fields are obtained. First of all, the stress components are related
to the stress function vector by

(55)
(56)

In the upper half-plane, the displacements are found to be

(57)

and
FRACTURE MECHANICS OF AN INTERFACE CRACK 199

(58)

where

(59)
(60)
(61)
(62)

(63)
(64)
(65)
(66)

(67)
(68)
(69)
(70)

(71)
(72)

(73)

(74)
200 L. BANKS-SILLS AND V. BONIFACE

B 1 = cos 2 + 12 sin 2 (75)


B 2 = cos2 + 22 sin2 . (76)
The in-plane stress components in the upper half-plane, denoted by the
superscript (1), are given by

(77)
FRACTURE MECHANICS OF AN INTERFACE CRACK 201

(78)

(79)

where

(80)
(81)
(82)
(83)
For the lower half-plane, the displacements are found to be
202 L. BANKS-SILLS AND V. BONIFACE

(84)

and

(85)

where
FRACTURE MECHANICS OF AN INTERFACE CRACK 203

N5 = cosh ( + ) (86)
N6 = sinh ( + ) (87)

M 9 = c o s (/2) (88)
M 10 = s i n (/2). (89)

The in-plane stress components in the lower half-plane are given by

(90)

(91)
204 L. BANKS-SILLS AND V. BONIFACE

(92)

where

M 11 = cos(3 /2) (93)


M 12 = sin(3 /2). (94)
PATH-INDEPENDENT INTEGRALS RELATED
TO THE J-INTEGRAL AND THEIR EVALUATIONS

SHIRO KUBO
Department of Mechanical Engineering and Systems
Graduate School of Engineering
Osaka University
2-1, Yamadaoka, Suita, Osaka, 565-0871 JAPAN

Abstract. In relation with Rices J-integral, local J vector, local J scalar,


global J vector and global J scalar proposed by the present author are in-
troduced, and their interrelations are discussed. It is shown that the local
J vector is perpendicular to the crack front. The global J scalar is inter-
preted in terms of the energy release rate with crack extension. The relation
between the global J scalar and the J-, M-, L-integrals are discussed, and
extended expressions of the M- and L-integral are presented. The local and
global J-integrals are applied to deduce J expressions for an axi-symmetric
crack. A method for estimating the distribution of J-integral along crack
front of a three-dimensional crack is described. Formulae for experimen-
tal evaluation of the J-integral and the modified J-integral for creep crack
problems using displacement measurement are introduced. Analytical ex-
pressions for several specific cracks are described.

1 . Introduction

The J-integral proposed by Rice (1968) has been playing a major role for
evaluating the behavior of cracks under elastic-plastic conditions. The line
integral expression of the J-integral was advantageously used in its numer-
ical evaluation. For experimental evaluation of the J-integral from load-
displacement curve, simple formulae proposed by Rice et al. (1973) pro-
moted extensively the application of the J-integral.
In this paper some integrals related to Rices J-integral and their interre-
lations are discussed. Formulae for experimental evaluation of the J-integral
and the modified J-integral (C * -integral) for creep crack problems (Ohji et
205
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 205221.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
206 S. KUBO

al., 1974, 1976; Landes and Begley, 1976) using displacement measurement
are introduced. Analytical expressions deduced by the present author for
several specific cracks are described.

2. Definition of Local and Global J-Integrals for a Three-Dimen-


sional Crack
Consider a cracked three-dimensional body in the Cartesian coordinates
x1 x2 x3 . The surface and the front of the crack are supposed to be smooth.
Assume the existence of strain energy density W, which relates strain i j
and stress i j as,
(1)

The definition of infinitesimal strain is used:

(2)

where u i is displacement and j following comma denotes the partial differ-


entiation with respect to x j . The traction Ti is defined in terms of j i and
the outward unit normal vector v j a s ,

(3)

The summation convention is used over the repeated index j. The crack
surface is supposed to be traction free. No body force is applied in the body,
and therefore the equilibrium equation is written as,

(4)

Local and global J vectors and scalars were proposed by Kubo et al.
(1981, 1982).

2.1. LOCAL J VECTOR


After Blackburn(1972) and Miyamoto et al. (1979) the local J-integral can
be defined at point O on crack front as follows, by taking the energy balance
of a plate of thickness B , which contains point O and is surrounded by
surfaces S 1 to S 5 shown in Fig. 1:

(5)

Here F is defined by
INTEGRALS RELATED TO THE J-INTEGRAL 207

Figure 1. Integration surface and path for local J-integral

(6)

where
(7)
The integral is evaluated for the surface of the plate, and i i denotes the
unit normal vector in the xi direction. The local J-integral J l o c al has three
components and can be defined as a vector.
It is easily shown that the local J-integral Jlocal is independent of the
selection of surfaces used for evaluation, using the definition of strain energy
density (Eq. (1)) and strain-displacement relationship (Eq. (2)) as is done
for proving the path-independence of the J-integral (Rice, 1968).
When the x 3 direction is taken to be in the thickness direction of the
plate used for the energy balance, the local J-integral Jlocal defined by Eq.
(5) can be rewritten as,

(8)

where is the contour of S 2 and is a path starting from the lower surface
of the crack and arriving at the upper surface in the x1 x 2 plane, and A i s
the surface surrounded by and the crack in the plane.
When curvilinear coordinates 1 2 3 are used, F is given by

(9)
208 S. KUBO

Figure 2. Integration surface for global J-integral

where g i is the contravariant base vector, and | i denotes the covariant


derivative.

2.2. LOCAL J SCALAR


The local J scalar can be defined as the absolute value of the local J vector:

J local = | J l o c a l |. (10)

2.3. GLOBAL J VECTOR

The global J vector for the Cartesian coordinates x1 x2 x 3 can be defined


by taking the integral of H over the surface S = S 1 + S 2 + S 3 + S 4 + S 5
surrounding the entire crack front as shown in Fig. 2.

(11)

When curvilinear coordinates 1 2 3 are used, J g l o b a l is given by

(12)

2.4. GLOBAL J SCALAR


Introduce a vector field m = mi i i , which is differentiable and is not singular
in the body. The global J scalar is defined by the following integral.

(13)
INTEGRALS RELATED TO THE J-INTEGRAL 209

Here V denotes the region surrounded by S and crack surfaces. It is easily


shown that J global (m) is independent of the selection of surface S.
When curvilinear coordinates are used, J global (m) is given by,

(14)

3. Direction of Local J Vector


When the x 3 direction is taken to be parallel to the tangential direction at
point O on the crack front, the x 3 component of the local J vector J l o c a l is
given as follows using the surfaces S 4 and S 5 defined in Fig. 1.

(15)
Denoting the distance from crack tip in the plate by r, i j ij has singularity
of the order of r 1. Stress ij has singularity of the order of r 1 / 2 or weaker
one and displacement u j has no singularity. The differentiation of these
values with respect to x 3 does not change the order of singularity. These
mean that the integrand of the first term on the right hand side of Eq. (15)
has the singularity of the order of r 1 . Then this integral approaches 0 when
integral domain A shrinks to O. The second term also approaches 0 when
A shrinks. Since J l o c a l is independent of the selection of integral domain A,
(J l o c a l)3 is always equal to zero. J l o c a l is thus perpendicular to the crack
front at each point of the crack front.

4. Interrelations Between J Scalars and Vectors


4.1. INTERRELATION BETWEEN LOCAL J VECTOR AND GLOBAL J
VECTOR
Consider layers of integral surfaces to cover a part of crack front shown in
Fig. 3. On surfaces facing with each other, outward unit normal vector v i is
opposite in direction, and therefore the contributions for these surfaces to F
are cancelled with each other. The sum of integration of H over the surface
of each layer is then equal to the integration of H over the cylindrical surface
covering the crack front . From Eqs. (5) and (11) the following equation
is thus obtained in the limit that the thickness B of the layer approaches 0.

(16)
210 S. KUBO

Figure 3. Layers of integration surfaces

Figure 4. Integration surface very close to crack front

In this equation the integration is taken along the crack front .

4.2. INTERRELATION BETWEEN LOCAL J VECTOR AND GLOBAL J


SCALAR
Take the integration surface of J global (m) very close to the crack front as
shown in Fig. 4. Since m has no singularity it can be taken as to be constant
for an infinitesimally small amount of crack front df.
Then the first term of the right hand side of Eq. (13) can be rewritten
as
(17)

where S is a path taken near the crack tip. The integrand of second term
of the right hand side of Eq. (13) has the singularity of the order of r 1 or
weaker one, and therefore the term approaches to 0 as the integration sur-
face S shrinks to the crack front. Therefore, the following equation relating
the global J scalar with the local J vector is obtained.

(18)
INTEGRALS RELATED TO THE J-INTEGRAL 211

4.3. INTERRELATION BETWEEN LOCAL J SCALAR AND GLOBAL J


SCALAR
When the angle between J l o c a l and m at each point of crack front is denoted
by , Eq. (18) is reduced to

(19)

where m denotes the absolute value of m.

4.4. INTERRELATION BETWEEN GLOBAL J VECTOR AND GLOBAL J


SCALAR
When vector m is constant along the crack front, Eq. (18) is reduced to,

(20)

Then using Eq. (16) the following relation is obtained.

Jg l o b a l (m) = m Jg l o b a l . (21)

5. Relation between Energy Release Rate and Global J Scalar


The potential energy II of the cracked body is given as,

(22)

where V T denotes the volume of the body, and U is the potential of external
forces.
Consider a sharp notch shown in Fig. 5. Suppose that the notch ex-
tended by m q and the volume of the body is reduced by V, where q
is a scalar independent of location. Then the change in the potential energy
is given by,
(23)

As was discussed by Budianski and Rice (1973) the first and the second
terms cancel with each other due to the principle of virtual work. The
change in the volume is given as,

d V = dS vi m i q, (24)
212 S.KUBO

Figure 5. Three-dimensional notch and its extension

where dS is a small element of notch surface SB . Then Eq. (23) is reduced


to,
(25)

When S B is used for the integral surface for Jglobal (m), Eq. (13) gives,

(26)

Since the notch is stress free T j = 0,

(27)

Comparing Eq. (25) with Eq. (27) and noting that the directions of v i
in these equations are opposite with each other, the following equation is
obtained.
J global ( m) = dII/ dq (28)
This equation states that J global ( m ) gives the release rate of the potential
energy when notch extends by m at each point of the notch tip.

6. Relation Between Global J Scalar and J-, M- and L-Integrals


6.1. RELATION BETWEEN GLOBAL J SCALAR AND RICES
J-INTEGRAL

Suppose that the vector m is equal to a constant c:

(29)

Then the global J scalar Jglobal ( m) = J global ( c ) is given as,

(30)
INTEGRALS RELATED TO THE J-INTEGRAL 213

This equation is rewritten as,

Jglobal (c) = c J global (31)

For a two-dimensional crack with

c1 = 1, c 2 = 0 , c 3 = 0, (32)

J global (c) is reduced to Rices J-integral.

6.2. RELATION BETWEEN GLOBAL J SCALAR AND M-INTEGRAL

When the vector m is given as,

(33)

the global J scalar Jglobal (m) = J global (x) is expressed as,

(34)

where i j denotes the Kronecker delta. For two-dimensional problems ii =


2 and for three-dimensional problems i i = 3.
For nonlinear materials whose strain is proportional to the n-th power
of stress, the strain energy density W is given by,

(35)

Then the integral over V in Eq. (34) is rewritten in terms of an integral


over S using the divergence theorem, and J global (x) is reduced to,

(36)

Consider a linear elastic material for which n = 1. The integration in


Eq. (36) is reduced to a line integral as follows for a two-dimensional crack
using ii = 2.
(37)

For a three-dimensional crack, i i = 3 and then Eq. (36) is reduced to,

(38)
214 S. KUBO

Equations (37) and (38) coincide with the expressions of the M-integral pro-
posed by Knowles and Sternberg (1972). Equation (36) gives an expression
of the M-integral extended for nonlinear materials.
When cylindrical coordinates r z are introduced, the components of m
are written as,
m1 = r, m 2 = 0, m3 = z , (39)
the global J scalar J global (m) = J global (r + z) is written as,

(40)

When spherical coordinates r are introduced, the components of m


are written as,
m 1 = r, m 2 = 0, m 3 = 0, (41)
and an expression of the global J scalar Jglobal ( m) = J global (r) in the spher-
ical coordinates is written as,

(42)

6.3. RELATION BETWEEN GLOBAL J SCALAR AND L-INTEGRAL

When the vector m is given as,

(43)

the global J scalar J global (e3 ik x k i i ) is expressed as,

(44)

using the divergence theorem, and the equilibrium equation (Eq. (4)) with
the permutation symbol e ijk .
When the cylindrical coordinates r z are used, and the components of
m are written as,

m 1 = 0 , m 2 = 1, m 3 = 0, (45)
INTEGRALS RELATED TO THE J-INTEGRAL 215

the global J scalar J global is written as,

(46)

When the spherical coordinates r are introduced, the components of


m are written as,
m1 = 0 , m 2 =0, m 3 = 1, (47)
an expression of the global J scalar J global in the spherical coordinates is
written as,

(48)

7. J-Integral for an Axi-Symmetric Crack


As a simplest example of three-dimensional cracks, an axi-symmetric crack
with radius a is discussed in this section.

7.1. J-INTEGRAL EXPRESSION BASED ON LOCAL J VECTOR

When the Cartesian coordinates are used, the local J vector at point O
shown in Fig. 6 is in i 1 direction and is given by,

(49)

since the local J vector is perpendicular to the crack front as is discussed


in section 3, and due to symmetry with respect to the crack plane.
Rewriting the components of stress and displacement in terms of the
cylindrical coordinates local J scalar Jlocal , which is equal to the magnitude
of J local , is given by,

(50)

7.2. J-INTEGRAL EXPRESSION BASED ON GLOBAL J VECTOR

When the global J vector is evaluated for surface S 1 + S 2 in Fig. 7,

(51)
216 S. KUBO

Figure 6. An axi-symmetric crack

Figure 7. An axi-symmetric crack cut into two halves along a symmetry plane

Since local J vector is constant in magnitude and is perpendicular to the


crack front, the global J vector, which is equal to the integration of the
local J vector along the crack front, is given as,

J global = 2i2 a Jlocal . (52)


From Eq. (51) the global J vector is given by the following equation for
the cylindrical coordinated r z.

(53)
INTEGRALS RELATED TO THE J-INTEGRAL 217

From Eqs. (52) and (53) the local J scalar J l o c a l is given by,

(54)

This expression multiplied by 2 a coincides with that proposed by Astiz et


al. (1975).

7.3. J-INTEGRAL EXPRESSION BASED ON GLOBAL J SCALAR

Suppose that the components of vector m is given as,

m 1 = (r), m2 = m3 = 0 (55)

in the cylindrical coordinates r z.

(56)

From Eq. (19) relating the global J scalar and local J scalar,

Jg l o b a l = 2a ( a ) J l o c a l . (57)

Then the local J scalar is expressed as,

(58)

Equation (58) is reduced to Eq. (54) and Eq. (50) for (r) = 1 and
(r) = 1/r, respectively. Equation (58) is a general expression of J-integral
for an axi-symmetric crack including Eqs. (54) and (50) as special cases.
Other expression can be obtained by employing other distributions of
m. The equivalence of the J value for various expressions was confirmed by
using the finite element analyses (Ohji et al., 1982). Since the value of local
218 S. KUBO

Figure 8. A semi-circular surface crack

J scalar is independent of the selection of the expression used, expressions


convenient for evaluation can be advantageously used.
When Eq. (39) is used in the cylindrical coordinates r z the local J
scalar for the n -th power nonlinear material can be reduced to,

(59)

8 . Estimation of Distribution of J-Integral Using Global J Scalar

The value of global J scalar J g l o b a l (m) can be obtained for various vector
m. By combining these values the distribution of local J can be estimated.
As an example, consider a semi-circular surface crack shown in Fig. 8.
Cylindrical coordinates r z are used. Since the local J vector is perpendic-
ular to the crack front and the problem is symmetrical with respect to the
crack plane, local J vector J l o c a l is in the r d i r e c t i o n .

Jl o c a l = J l o c a l g1 , (60)
where g1 denotes the contravariant base vector in the r direction. The
distribution of J l o c a l can be expressed using the Fourier series as,

(61)
Take the form of the vector m expressed as,

m k = c o s k g1, (62)
where g 1 is the covariant base vector in the r direction. Then from Eqs.
(18), (60), (61) and (62)

Jg l o b a l ( m k ) = ab k /2. (63)
INTEGRALS RELATED TO THE J-INTEGRAL 219

J global ( m k ) is given by introducing m k in Eq. (14). Then from Eq. (63),


the value of b k determining the distribution of J, is given by,

(64)

The global J scalar method was applied to calculate the distribution


of J-integral along crack front of a three-dimensional crack using the finite
element stress strain analysis (Ohji et al., 1986a).

9. Simple Methods for Experimental Determination of J-Integral


and Modified J-Integral
Rice et al. (1973) proposed simple formulae for determining J values exper-
imentally using load-displacement curve. Some extensions of the method
were made (Kubo, 1975, Ohji et al., 1978).
The method was successfully extended (Kubo, 1975, Ohji et al., 1978)
for determining the modified J-integral J * (Ohji et al., 1974, 1976), which
is equivalent to the C * integral (Landes and Begley, 1976), for creep crack
problems. For example J* value for a plate with deep edge cracks can be
given as,
(65)

where n is the creep exponent, net is the net section stress, and is the
load-point displacement rate. When opening rate n of crack edge is used
in place of , the J* value for shallow cracks as well as for deep cracks can
be well approximated:
(66)

10. Simple Formulae for Analytical Estimation of J-Integral


Ohji et al. (1981) proposed simple formulae for estimating J values for a
small crack in large plate. It was assumed that strain is proportional to
the n -th power of stress. For mode III crack of size 2a, J value can be
220 S. KUBO

approximated by,

(67)

where net is the net section shear stress, and net is the corresponding shear
strain. For mode III deep crack of ligament size b, J value is approximated
by,
(68)
For a mode III center crack in a plate with crack length of 2a and ligament
length of b, approximate J value can be obtained by combining Js and J d
values given by Eqs. (67) and (68).

(69)

This equation is in analogy with Neubers rule for notch stress concentra-
tion.
Similarly, for mode I shallow crack, J value is approximated by,

(70)

where net is the net section tensile stress, and net is the corresponding
strain. For mode I deep crack of ligament size b, J value is approximated
by,
(71)

For a mode I center crack in a plate with crack length of 2a and ligament
length of b, approximate J value can be obtained by applying Eq. (69) using
J s and J d values given in Eqs. (70) and (71).
For a mode III crack of length a emanating from a notch of radius , J
value is well approximated by the following equation (Ohji et al., 1986b).

(72)

Here max and max are the maximum shear stress and shear strain at notch
root in the absence of the crack, and k being a constant dependent on the
strain hardening exponent n. For n = 1, k = 1.5 and for n = 9, k 3

Acknowledgements
The work is partly supported by the Ministry of Education, Science, Sports
and Culture under Grant-in-Aid for Scientific Research.
INTEGRALS RELATED TO THE J-INTEGRAL 221

References

Astiz, M.A., Elices, M., and Galvez, V.S. (1975) Proc. ICF4, Waterloo, Canada
3, 395.
Blackburn, W.S. (1972) Int. J. Fracture 8, 343.
Budianski, B., and Rice, J.R. (1973) J. Appl. Mech. 40, 201.
Knowles, J.K., and Sternberg, E. (1972) Archive for Rat. Mech. Anal. 44, 187.
Kubo, S. (1975) Study on Mechanics of Crack Initiation and Growth, Osaka
University Doctoral Thesis (in Japanese)
Kubo, S., and Ohji, K. (1981) J. Soc. Mater. Sci., Japan 30, 796 (in Japanese).
Kubo, S., and Ohji, K. (1982) J. Soc. Mater. Sci., Japan 31, 32 (in Japanese).
Landes, J.D., and Begley, J.A. (1976) Mechanics of Crack Growth, ASTM STP
590, 128.
Miyamoto, H., and Kikuchi, M. (1979) U.S.- Japan Seminar on Fracture Mechan-
ics of Ductile and Tough Materials and Its Application to Energy Related
Structures, Hayama, Japan, 33.
Ohji, K., Ogura, K., and Kubo, S., (1974) Preprint Japan Soc. Mech. Engrs. No.
740-11, 207 (in Japanese).
Ohji, K., Ogura, K., and Kubo, S. (1976) Trans. Japan Soc. Mech. Engrs. 42, 350
(in Japanese).
Ohji, K., Ogura, K., and Kubo, S. (1978) Trans. Japan Soc. Mech. Engrs. 44,
1831 (in Japanese).
Ohji, K., Ogura, K., and Kubo, S. (1981) Trans. Japan Soc. Mech. Engrs., Ser.
A 47, 400 (in Japanese).
Ohji, K., Kubo, S., and Suehiro, S. (1982) Preprint Japan Soc. Mech. Engrs. No.
820-2, 57 (in Japanese).
Ohji, K., Kubo, S., Suehiro, S., and Nishimura, K. (1986a) Trans. Japan Soc.
Mech. Engrs., Ser. A 52, 1034 (in Japanese).
Ohji, K., Kubo, S., Maeda, T. (1986b) Trans. Japan Soc. Mech. Engrs., Ser. A
52, 1300 (in Japanese).
Rice, J.R. (1968) J. Appl. Mech. 35, 379.
Rice, J.R., Paris, P.C., and Merkle, J.G. (1973) Progress in Flaw Growth and
Fracture Toughness Testing, ASTM STP 536, 231.
This page intentionally left blank.
ON THE EXTENSION OF THE JR CONCEPT TO SIGNIFICANT
CRACK GROWTH

B. COTTERELL AND Z. CHEN


Institute of Materials Research and Engineering
3 Research Link, Singapore 117602

AND

A. G. ATKINS
Department of Engineering
University of Reading
Reading, BG6 6AY, UK

Abstract: A possible method of estimating the extra work required in real elasto-plastic
fracture as compared with a hypothetical non-linear elastic specimen is discussed. It is
shown how the extra work term can be used to calculate the specific essential work, R.,
from J R obtained by the standard ASTM method. Although the method presented
apparently underestimates the extra work, it points a way forward to obtain the true crack
growth resistance.

1. Introduction

The two most accepted fracture mechanics concepts are the stress intensity factor for
linear elastic materials established by Irwin (1957) and the J-integral established by
Rice (1968) for elasto-plastic materials. As with any concept, both have been at times
pushed beyond their regimes of validity. The J-integral is exact for a non-linear elastic
(nle) material, but can be applied in many cases to an elasto-plastic (elp) material. In an
elp material, the value of J-integral at initiation is equal to the specific essential work,
R, performed within the fracture process zone (FPZ) at the tip of a crack. Provided the
FPZ is small, J can be obtained from the potential energy or the non-linear strain energy
release rate and is given by

(1)

where II is the potential energy of the system, nle , is the non-linear strain energy, a is
the crack length, B is the thickness of the specimen, and u is the displacement. In
practice J is usually obtained from this energetic interpretation rather than from the
original contour integral.
223
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 223235.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
224 B. COTTERELL, ET AL.

Figure 1. Schematic load-displacement curves for a real elp and hypothetical nle
material

Schematic load-displacement curves for a real elp and a hypothetical nle specimen are
shown in Figure 1. The nle specimen is assumed to be made from a hypothetical nle
material that has the same stress-strain curve as the elp material during loading but
which is completely elastic, and has the same specific work of fracture, R, which, if not
a constant, has the same dependence on the crack extension, a. Up to the initiation of
a crack from a notch of length ai , the elasto-plastic deformation closely follows
Henckys equations and is practically identical to the hypothetical nle curve. At I a
ductile crack is initiated in both the elp and nle specimens. After initiation the load paths
of the two specimens diverge. Depending upon the material and the specimen geometry,
the load can either increase or decrease during propagation. Some elastic energy
remains locked into the specimen on unloading because of the non-uniform plastic
deformation and it is simplest to divide the work done up to the initiation of a crack
(area OID) into a non-recoverable work, i (area OIE), and a recoverable elastic strain
energy, (area IDE). The non-recoverable work, i , is the plastic work combined with
the non-recoverable elastic strain energy. The initiation value, Ji , of the J-integral is
equal to the initiation value of the specific essential work of fracture, Ri . Often J is
divided into a plastic part, Jp , and an elastic part, At crack initiation the plastic
component, Jpi , can be obtained from i and is given by (Sumpter & Turner 1976)

(2)

where is a factor that depends on the geometry, and b i is the remaining ligament at
initiation. For the deep notch bend specimen =2; for other geometries , though not
a constant, only weakly depends on the crack length.
During crack propagation non-proportional unloading occurs behind the crack tip
in a elp specimen, which cannot be modelled by a nle theory except under certain
restrictive conditions. Hutchinson and Paris (1979) have shown that outside of the
region of non-proportional loading, there is a region where the deformation is nearly
proportional. If this region is well contained within the region dominated by the J-
J R CONCEPT TO CRACK GROWTH 225

singularity, there exists an annular region where the Hutchinson-Rice-Rosengren


equations hold (Hutchinson 1968; Rice and Rosengren 1968). Limitations on the slope
of the crack growth resistance, as assessed by J, have been developed to try to ensure
that crack growth is controlled by the J-integral (McMeeking & Parks 1979; Shih &
German 1981; Hutchinson 1983). A further complicating factor is the dependence of
even Ji on the degree of constraint which can only be accommodated by introducing a
second parameter as for example the Q-stress suggested by ODowd and Shih
(1991,1992). This paper will concentrate on the inconsistencies introduced by using a
nle theory for propagation. Unless the plastic zone is well contained within an elastic
stress field, the value of the propagation J-integral, JR , as determined by the standard
method (ASTM 1987), is only independent of size and geometry for small crack
extensions (Hancock et al. 1993; Joyce & Link 1995; Xia et al. 1995; Xia & Shih 1995).
The philosophy for calculation of the plastic component, Jp , during crack
propagation is to assume that the deformation along the path IA in Figure 1 is nle
(Hutchinson and Paris 1979; Ernst et al. 1981; ASTM 1987). This feature, analyzing
crack propagation in a real elp specimen as if it was nle, is one of the keys allows us to
unlock the specific work of fracture, R, from JR calculated by the standard ASTM
method. A simplified version of the integral equation formulated by Ersnt et al. (1981)
can be obtained, if it is assumed that is a constant, by the differentiation of Eq. (2) and
is

(3)

where X is the load and u p is the plastic part of the displacement. In the Ernst et al.
(1981) and the ASTM (1987) Standard methods, Eq. (3) is integrated numerically along
the propagation path IA in Figure 1. However, the real elp displacements are larger than
the equivalent nle displacements by an amount u extra and the load, X nle , on the equivalent
nle load-displacement curve is not identical to the real elp load, X e l p , except for the
special case of the double cantilever beam (DCB) geometry. The Ernst et al. (1981)
method overestimates Jp by including some of the plastic work outside of the FPZ
making many JR curves spurious (Cotterell & Atkins 1996). Thus the so-called crack
resistance is often an artifact of the method of calculating JR . True crack growth
resistance as assessed by the specific essential work probably can arise only from only
two causes: shear lip formation, the original cause proposed for crack growth resistance
(Kraft et al. 1961), and changes in the constraint which can be analysed by such methods
as the strip FPZ model based upon a Gurson material (Xia & Shih 1995; Xia et al. 1995).
Experimental evidence for the above statements has been given by Atkins et al. (1998)
and is briefly summarised here.
The crack propagation in one very special test geometry, the deep side-grooved double
cantilever beam (DCB) specimen loaded by an end moment, is steady state from
initiation and thus there should be no crack growth resistance because shear lip
formation is suppressed by the side grooves and there is no change in constraint at the
crack tip. Hence the specific essential work of fracture, R, must be a constant. Provided
the beam crack length to height ratio is reasonably large, the crack propagation in an
end-force-loaded DCB specimen is close to steady state. The load-displacement curves
for two aluminum alloy DCB specimens, one with a notch length of 70mm with a crack
propagated to a total length of 140mm and one whose notch length is 140mm, are shown
226 B. COTTERELL, ET AL.

in Figure 2. The initiation load for the specimen with the long notch and the load after
propagation to the same total crack length for the short notch are identical, but the beam
displacement for the second case is much larger. If the aluminum alloy were nle the load-
displacement point for both initiation from the long notch and propagation to the same
total crack length from the short notch would be identical. In this case J calculated for
both specimens would be the same and equal to the specific essential work, R. However,
when the specimen with the short notch is analyzed using the Ernst et al. (1981)
technique, the resulting JR -curve indicates a spurious crack growth resistance (Atkins et
al. 1998).

Figure 2. Load-displacement curves for two elp aluminium specimens


(After Atkins et al. 1998)

The DCB specimen is peculiar because the crack propagates into virgin material.
The moment at the crack tip remains constant during crack propagation. Thus in an end-
force-loaded DCB specimen, the force for crack initiation is identical to the force
necessary to propagate a crack to the same length but, as seen in Figure 2, the
displacement in the propagated crack is larger by an amount which we call uextra . The
extra work performed by a epl specimen propagated from a notch length a i to a crack
length a over that performed on a hypothetical nle specimen is given by the areas
IAABBI in Figures 1 and 2. The extra work is the nle work that is not recovered in the
case of elp. In this paper we discuss a possible way of back calculating from a real elp
load-displacement curve, the corresponding curve for a hypothetical nle specimen that
does not include the extra work term. This virtual nle curve can then be analysed by the
Ersnt et al. (1981) method to obtain the true value of JR which is the specific essential
work of fracture, R.
Chen (1997) has shown that, in the special case of the DCB specimen, it is possible
to calculate the elp curve directly from the nle curve. Because of the peculiarity of the
DCB specimen, the recoverable elastic energy in an elp specimen is identical to the
recoverable linear elastic energy in the hypothetical nle specimen of the same crack
length. The rate of external work in a elp specimen and in the hypothetical nle one are
given by
J R CONCEPT TO CRACK GROWTH 227

(4)

where w elp is the plastic work performed to create a unit crack extension, and nle is the
non-linear part of the elastic strain energy. During propagation the bending moment is
crack initiation. Thus during propagation

(5)

In the hypothetical nle specimen non-linear elastic energy stored behind the crack tip
is recovered as well as the linear elastic energy and hence

(6)

The displacement u elp can therefore be found exactly from u n le by integration of Eq


(6) for the DCB specimen (Chen 1997). The question is can a similar technique be
used on standard test geometries?

2. A Posssible Link between ELP and NLE

We have shown how the scheme proposed by Ernst et al. (1981) includes in JR some of
the plastic work performed outside of the FPZ. Here we discuss an approximate method
for extracting R from conventional JR -curves obtained from specimens that are fully
yielded at initiation. Consider first a nle specimen made from a hypothetical material
defined in Section 1. Ahead of the crack tip there is a region of area S where the stresses
are plastic (see Figure 3). Assume for the moment that the plastic energy density
within this zone is uniform and equal to p so that n le = S p . As the crack propagates
both the size of the plastic region and the plastic energy density change. The rate
of increase in plastic energy is given by

(7)

the first term in Eq. (7) comes from the change in plastic energy density and the

For a nle material the plastic region is defined as the region where the stress-strain
relationship is non-linear.
228 B. COTTERELL, ET AL.

second term is numerically equal to the plastic energy recovered as the plastic zone
shrinks since the material is nle. In an elp material the plastic work cannot be recovered
and it is argued that this second term is numerically equal to the extra work that must be
performed in a real elp specimen as compared with the hypothetical nle specimen. Thus
for an elp specimen

(8)

where Uextra is the extra work performed in the real elp specimen. The extra work,
assuming the area of the active plastic region is proportional to b , is given by

(9)

The plastic energy (or work) density for the real elp specimen cannot be exactly the
same as that for the hypothetical nle specimen, but we argue that, since the active plastic
region is one where the material is being loaded, the plastic energy will be approximately
the same. Also critical is the assumption that the plastic energy density is uniform.
Clearly the plastic energy density cannot be uniform at initiation because by definition
at the boundary of the plastic region the stresses must be elastic and the plastic energy
density zero. However, it is argued that if the plastic zone at the tip of the propagating
crack is contained within the plastic zone at initiation then the plastic energy density
becomes more uniform with crack growth and the accumulated error in the method is
likely to be small. Figure 3 shows schematically how the plastic energy density develops
with crack growth from an initial notch at position 1 through to a crack at position 3. The
development of the plastic energy density along a line X-X is shown at the bottom of the
figure. At initiation the plastic energy density is zero at the edge of the plastic zone.
However, if the boundary of the active plastic zone is nested within the previous plastic
zone, then the plastic energy density along the boundary will accumulate with crack
growth and the plastic energy density along the active portion of the line X-X will
become more uniform. It is also assumed that the shape of the plastic zone remains
similar during crack propagation so that its area is proportional to b .
It is our intention to show how the specific essential work of fracture, R, may be
obtained approximately from J R using the concept of extra work. For the moment
consider a hypothetical nle specimen and assume that, up to initiation its load-
displacement curve is identical to the real elp specimen. After the crack grown to a crack
length a (point A in Figure 4), any unloading will be along AO. The plastic
component of the specific essential work Rp is given by

(10)
JR CONCEPT TO CRACK GROWTH 229

Figure 3. Schematic development of Figure 4. Schematic load-displacement curve


plastic region and energy density

where nle =area OA'C'O. Since the fracture is nle, no energy is dissipated except at the
crack tip and hence the area OIA'O is given by

(11)

The accumulated plastic work of the hypothetical nle specimen (area OIA'C'O) can be
obtained from Eqs. (10) and (11) and is

(12)

From Eq. (9) the extra work , U extra , performed by the real elp specimen is given by

(13)

Hence the accumulated plastic work of the real elp specimen , defined by the area
OIACO, is the sum of Eqs. (12) and (13)

(14)

Now our argument is that since J p is obtained in the standard ASTM method by
pretending that the real elp specimen behaves as if it were nle, the accumulated work is
also given by
230 B. COTTERELL, ET AL.

(15)

Hence the differential equation connecting Jp and R p i s

(16)

If is a constant Eq. (16) can be rewritten as

(17)

Note that Eqs. (16) and (17) do not depend on any modelling of the load-displacement
curves for either the true elp specimen or its equivalent nle specimen. We now explore
some of the implications of the extra work concept.
First let us assume that the elastic components of R and J are very small and look at the
deep notch bend specimen where =2. If the specific essential work of fracture, R, is
constant, Eq. (17) implies that

(18)

where a is the crack extension. A size effect is apparent. However, most JR -curves are
not linear. In the Section 3 we examine the effect of elastic deformation.

3. Elasto-plastic Crack Growth

As the size of a specimen increases so does the proportion of elastic deformation. Hence
specimen size has a large effect on the JR -curve. For the purposes of illustration we
examine the behavior of an elasto-plastic deep-notch three point bend specimen with a
notch ai /W=0.5 and assume that the load, X, in both the nle and elp specimens is given
by

(19)

The EPRI analyses use this separable variables power law (EPRI 1984). The exponent,
n, of the plastic displacement, u p , can be identified loosely with the strain hardening
exponent, and the squared on ligament, b, is consistent with the -factor of 2 for deep
notch bend specimens. The elastic displacement is obtained from the compliance,
C(a/w), given in the ASTM Standard Method for Determining J-R Curves (ASTM
E1 152-1987) so that
J R CONCEPT TO CRACK GROWTH 231

(20)

where S is the span and E the elastic modulus. The ratio S/W has been taken as 4.
The inverse of the real problem has been examined, that is a specific essential crack
growth resistance, R( a), is assumed and the expected JR -curve calculated. However, in
principle the real problem of calculating R from JR is no different. At initiation Ji=R i and
JR are normalized by Ri . Dimensional analysis shows that the normalized JR d e p e n d s
upon 0W/R i ,0 /E, R/Ri , and n . In this simple illustrative exploration a typical
value of 0.1 has been chosen for n.

3 . 1 J R -CURVES FOR A CONSTANT ESSENTIAL FRACTURE WORK

The specific essential work of fracture has an elastic, Re , as well as a plastic component,
Rp . Using Eq.(9) and adding an elastic term, R is given by

(21)

where G e (a/W) has been obtained from the expression for the stress intensity factor given
in ASTM E1 152-1987. The difference between the plane stress and plane strain elastic
modulus has been ignored. For any crack length this equation has been solved to give
u nle /W which then has been used to calculate the components R p and R e . Eq. (17) has then
been integrated to give Jp . Making use of the fact that the ASTM method tacitly assumes
nle behavior, u elp is given by

(22)

The elastic component J e and the elastic displacement have then been calculated from
the load. JR and load-displacement curves are shown in Figures 5 & 6 for a range of
specimen sizes based on a unit specimen where 0 W0 /Ri =10, 0 /E=0.0005. For small
specimens the elastic strains are very small and JR is a function only of a/W and is given

Figure 5. JR versus a/W curves Figure 6. Load-displacement curves


232 B. COTTERELL, ET AL.

by Eq. (18). For larger specimens the crack growth resistance is less. In Figure 5,
specimens larger than 10 times the unit size do not yield completely before initiation and
are outside of the validity of the theory but the J R -curves have been included as they do
show qualitatively how J R is size dependent. The load-displacement curves show a
discontinuity in slope at initiation which is not observed in ductile specimens. However,
if the R-curve is given some crack growth resistance, load-displacement curves without
slope discontinuities can be obtained as is shown in the Section 3.2.

3.2 J R -CURVES FOR INHERENT CRACK GROWTH RESISTANCE

As discussed in the introduction there can be a real R crack growth resistance. In this
exploratory paper, this crack growth resistance is modeled by the exponential expression

(23)

The analysis follows that of Section 3.1, except that now instead of a constant on the left
hand side of Eq. (21) there is the expression contained in Eq. (23). In the JR and load-
displacement curves shown in Figures 7 and 8, the two parameters chosen to define R
in Eq. (23) are: R=4 and =5 the other parameters: 0 W0 /Ri =10, 0 /E=0.0005, and
n=0.1, are the same as in Figures 5&6.

Figure 7. J R and R-curves Figure 8. Load-displacement curves

Figure 7 shows both the crack growth resistance of the specific essential work of
fracture, R, and J R. For very small crack growths JR is almost identical to the specific
essential work of fracture, but the differences become marked for large crack growths.
The difference between J R and R reflects the extra plastic work performed in real elp
specimens over that in hypothetical nle specimens and does not reflect a real increase in
fracture resistance. When some real crack growth resistance in the specific essential
work of fracture is introduced, fracture initiation can occur prior to the attainment of the
maximum load without a discontinuity in the load-displacement slope (see Figure 8).
J R CONCEPT TO CRACK GROWTH 233

4. Comparison of Extra Work with Other Predictions

Figure 9. JR -curves for 3-point bend A533B specimens, a/W=0.6.

Xia et al. (1995) have shown that accurate predictions of load-displacement and JR -
curves can be obtained by embedding a FPZ, in the form of a row of void-containing cell
elements ahead of the crack, within a conventional elasto-plastic continuum. In particular
Xia et al. (1995) were successful in modeling the experimental JR -curves obtained by
Joyce & Link (1994) for A533B steel specimens of differing geometries. Xia et al.
(1995) also have predicted the JR -curves for deep notch bend specimens whose depth
ranges from 50 to 300 mm. There is complete yielding in specimens whose depth, W,
is up to 200 mm. However the largest specimen, W=300 mm, is globally elastic. What
we have done is to assume that the JR -curve for this largest specimen corresponds closely
to the specific essential work of fracture, R. Using this assumed R-curve, we have
predicted the JR -curves from Eq. (17), neglecting any elastic effects. These predicted JR -
curves are compared with those calculated by Xia et al. (1995) in Figure 9. Although the
general trend of increased J R is shown, the predicted values are very significantly
smaller.

5. Conclusions

We have shown that the difference between a real elp specimen and its equivalent
hypothetical nle counterpart, lies in an extra work term which is the energy that is not
recovered in a real specimen as the plastic zone contracts. Hence because in the standard
ASTM method J R is obtained by an analysis that assumes a nle deformation, it is possible
to extract the specific work of fracture, R, from JR . The difficulty lies in expressing what
is the plastic work accumulated in the active plastic zone.
In this paper the inverse problem has been has been analyzed. A specific essential work
234 B. COTTERELL, ET AL.

of fracture, R, has been assumed and JR that would have been obtained from the ASTM
standard method calculated. The general size effect on JR has been modeled, but the
extra work, and hence JR, has been underestimated. Further work to determine a simple
procedure to enable the specific essential work and the size dependence of JR to be
accurately estimated is worthwhile since much effort has been expended on obtaining
J R -curves that are size and geometry dependent.

References

ASTM E 1152-87 (1987) Standard Test Method for Determining JR curves


Atkins, A.G., Chen, Z. and Cotterell, B. (1998) The essential work of fracture and JR curves
for the double cantilever beam specimen: an examination of elasto-plastic crack propagation,
Proceedings of the Royal Society of London, A 454,815-833.
Chen, Z. (1997) Elasto-Plastic Fracture Propagation in Cantilever Beams, Doctoral Thesis,
University of Reading, Reading, UK.
Cotterell, B. and Atkins, A.G. (1996) A review of the J and I integrals, International Journal of
Fracture, 81, 357-372.
EPRI (1984) Advances in Elastic-Plastic Fracture Analysis, Electric Power Research Institute.
Ernst, H.A., Paris, P.C. and Landes, J.D. (1981) Estimations on the J-integral and tearing
modulus from a single specimen test record, in Fracture Mechanics: 13th Conference, ASTM
STP 743, R. Roberts (ed.), American Society for Testing and Materials, Philadelphia, 476-
502.
Hancock, J.W., Reuter, W.G. and Park, D.M. (1993) Constraint and toughness parameterized,
in Constraint Effects in Fracture, ASTM STP 1171, W.M. Hackett, K.H. Schwabe and R.H.
Dobbs (eds.), American Society for Testing and Materials, Philadelphia, 21-40.
Hutchinson, J.W. (1968) Singular behaviour at the end of a tensile crack in a hardening material,
Journal of the Mechanics and Physics of Solids, 16, 13-31.
Hutchinson, J.W. (1983) Fundamentals of the phenomenological theory of non-linear fracture
mechanics, Journal of Applied Mechanics, 35, 1042-1051.
Hutchinson, J.W. and Paris, P.C. (1979) in Elastic-Plastic Fracture, ASTM STP 668, J.D.
Landes, J.A. Begley, and G.A. Clarke (eds.), American Society for Testing and Materials,
Philadelphia, 37-51.
Irwin, G.R. (1957) Analysis of stresses and strains near the end of a crack traversing a plate,
Journal of Applied Mechanics, 24,361-364.
Joyce, J.A. and Link, R.E. (1995) Effects of constraint on upper shelf fracture toughness, in
Fracture Mechanics: 26th Volume, ASTM STP 1256, W.G. Reuter, J.H. Underwood, and
J.C. Newman (eds.), American Society for Testing and Materials, Philadelphia, 142-163.
Krafft, J.M., Sullivan, A.M. and Boyle, R.W. (1961) Effect of dimensions on fast fracture
instability of notched sheets, In Proceedings Symposium on Crack Propagation, College of
Aeronautics, Cranfield 8- 15.
McMeeking, R.M. and Parks, D.M. (1979) On criteria for J-dominance of crack fields in large
scale yielding, in Elastic-Plastic Fracture, ASTM STP 668, J.D. Landes, J.A. Begley, and
G.A. Clarke (eds.), American Society for Testing and Materials, Philadelphia, 175-194.
ODowd, N.P. and Shih, C.F. (1991) Family of crack-tip fields characterized by a triaxiality
parameter - I Structure of fields, Journal of the Mechanics and Physics of Solids, 39,989-
1015.
J R CONCEPT TO CRACK GROWTH 235

ODowd, N.P. and Shih, C.F. (1992) Family of crack-tip fields characterized by a triaxiality
parameter - II Fracture applications, Journal of the Mechanics and Physics of Solids, 40,939-
963.
Rice, J.R. (1968) Mathematical Analysis in the Mechanics of Fracture, in Fracture, H.Liebowitz
(ed.), Academic Press, New York Vol II, 192-314.
Rice, J.R. and Rosengren (1968) Plane-strain deformation near a crack tip in a power law
hardening materials, Journal of the Mechanics and Physics of Solids, 16, 1-12.
Shih, F.C. and German, M.D. (198 1) Requirements for a one parameter characterization of crack
tip fields by the HHR singularity, International Journal of Fracture, 17,27-43.
Sumpter, J. G. D. and Turner, C. E. (1976) Method for laboratory determination of JIC, in Cracks
and Fracture, ASTM STP 601, ASTM, Philadelphia, 3-18
Xia, L. and Shih, C.F. (1995) I. A numerical study using computational cells with
microstructurally based length scales, Journal of the Mechanics and Physics of Solids, 43,
233-259.
Xia, L., Shih, C.F. and Hutchinson, J.W. (1995) A computational approach to ductile crack
growth under large scale yielding conditions, Journal of the Mechanics and Physics of Solids,
43, 389-413.
This page intentionally left blank.
EFFECT OF T-STRESS ON EDGE DISLOCATION FORMATION
AT A CRACK TIP UNDER MODE I LOADING

G. E. BELTZ AND L. L. FISCHER


Department of Mechanical and Environmental Engineering
University of California
Santa Barbara, CA 93106-5070, USA

Abstract: We calculate the effect of the nonsingular stress acting parallel to a crack
(the T-stress) on edge dislocation nucleation at a crack loaded in Mode I. We find
that this leads to crack size effect that is, for small cracks (of order 100 atomic
spacings or less), the T stress causes the critical load for dislocation nucleation
(expressed in terms of the applied stress intensity factor) to deviate from the classical T
= 0 result. Specific results are discussed for the case of a finite crack subject to remote
tension, where it is shown that the threshold for dislocation nucleation is reduced.

1. Introduction

In the continuum modeling of atomic-scale phenomena at crack tips, an important


feature of the asymptotic representation of the stress field is often overlooked.
Following the work of Williams [1], the expansion of the stress field in cylindrical
coordinates about the tip (see Figure 1) may be generally written as

(1)

where K is the well-known applied stress intensity factor, and Sij ( ), Tij ( ), etc.
represent the angular variation of the field. The first term is singular in r, the second
term remains finite in the vicinity of the tip, and the remaining terms vanish as r 0 .
Linear elastic fracture mechanics is based on the reasonable notion that fracture
processes that occur close to the tip are only affected by the singular contribution;
hence, only the first term of Equation (1) is acknowledged as a valid descriptor of the
stress field in a vast majority of the fracture literature. The second term has a
particularly simple form that is, it can be shown [1] that T11 , T 33, and T13 (= T 31 ) must
be constant in in order to satisfy the field equations of elasticity, and the remaining
components must vanish in order to preserve the traction-free boundary condition on
the crack faces. We will exclusively deal with plane strain for the remainder of this
discussion, thereby rendering T11 the only component of interest. Then, the asymptotic
representation of the stress field around a crack, ignoring terms of r1/2 and higher, takes
the form
237
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 237242.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
238 G.E. BELTZ AND L.L. FISCHER

(2)

The constant term T11 = T is commonly known as the T-stress.


Several researchers have demonstrated that the T-stress has a significant influence
on the shape and size of the plastic zone that develops at a crack in a ductile material
[2-5] as well as the directional stability of an advancing crack in a brittle solid [5-8].
The latter effect can be qualitatively understood by imagining a slight upward
perturbation of the crack path (see Figure 1): if T > 0, the crack tip will continue to
undergo opening forces, and will continue to veer away from the x1-axis. If T < 0, the
crack opening forces tend to decrease, such that the only way for the crack to continue
to propagate is for it to remain on the x1 -axis.
Here, we will focus on the effect of T-stress on the threshold for dislocation
nucleation. The basic framework for analysis will be that due to Peierls [9] and
Nabarro [10], as introduced by Rice [11] for dislocation formation at a crack. We
directly calculate the critical stress intensity factor K crit for dislocation nucleation as a
function of T, and comment on the situations for which this effect could be most
significant.

2. Peierls Framework for Dislocation Nucleation at a Crack

The Peierls framework for dislocation formation at a crack assumes that the
dislocation/crack system can be thought of as two elastic semispaces separated by a
common plane (the crack plane and slip plane) on which there is a discontinuous jump
in the displacement fields [11]. There exists a periodic relationship between shear
stress and slip displacement along the slip plane, with traction free surfaces along the
crack plane, as shown schematically in Figure 1. Prior to dislocation nucleation, there

Figure 1. Schematic of crack and slip plane (dashed) inclined at angle . The T-stress gives a
contribution to 11 in addition to the classical K field result. A positive T-stress decreases the
resolved shear stress on the slip plane, while a negative T-stress increases the resolved shear
stress.

is a distribution of slip discontinuity along the slip plane that ultimately reaches a point
of instability with increased applied load and results in the nucleation of a dislocation.
T-STRESS EFFECT ON DISLOCATION FORMATION 239

Using this theory, the critical stress intensity factor for various fixed values of the T-
stress may be determined, thus establishing a locus of K and T values at which
nucleation occurs.
For all of the calculations presented here, a simple relationship due to Frenkel [12]
between shear stress and slip displacement on the slip plane is assumed,

(3)

where is the shear stress, is the relative atomic displacement between two atomic
planes, h is the interplanar spacing of those two planes, is the shear modulus, b is the
Burgers vector, and us is the unstable stacking energy (equal to b /2 h in the Frenkel
model). As introduced by Rice [11], the continuum analog to (referred to as ) is
thought of as extrapolated to a cut halfway between the slipping planes and is given
by

(4)

From elastic considerations, the stress along the slip plane can be written as:

(5)

where the first two terms on the right hand side give the pre-existing shear stress along
the slip plane due to the applied load on the crack geometry (comprising the most
singular term, scaled by K, as well as the constant term, proportional to T ), and the
third term reflects the stress relaxation that occurs due to sliding along the cut. The
kernel in the integral term represents the stress due to a dislocation positioned at s,
integrated over the entire slip distribution consisting of a continuous array of
infinitesimal Burgers vectors We seek a slip distribution (r) such that,
for all r > 0, [ (r)] predicted by the linear elastic formulation, Equation (5), must equal
[] provided by the atomic-based shear relation in Equation (3). Using a numerical
procedure outlined by Beltz [13] and Beltz and Rice [14], we carry this out for
incremental increases in K, at a fixed value of T, until an instability (i.e., dislocation
nucleation) is attained.

3. Results and Discussion

The principal results are shown in Figure 2, where the critical stress intensity factor for
dislocation nucleation is plotted as a function of T-stress for several slip plane angles
between 0 and 90. It has been shown in earlier work [11,15] that the parameter
2 u s /(l ) is the natural normalization factor for Kcrit , as it represents the
threshold for dislocation nucleation under the simplest geometry of a mode II shear
240 G.E. BELTZ AND L.L. FISCHER

Figure 2. Positive (or tensile) T-stress increases the critical stress intensity for nucleation, while a negative
T-stress decreases the critical stress intensity for nucleation.

crack with a non-inclined slip plane. The vertical axis is compressed logarithmically in
order to capture results for a wide range of slip plane inclination angles (ranging from
15 to 85). The general trend is that positive values of T increase the threshold for
dislocation nucleation, while negative values of T lower the threshold. The results for
T = 0 are consistent with earlier work by Rice et al. [15]. For 0, the critical load
would of course become unbounded since the resolved shear stress driving dislocation
formation would vanish. The T-stress effect is greatest for intermediate angles, and
can cause variations in the critical load by up to approximately 50% for the T values
considered here, depending on the sign of the T-stress. We note that the T-stress effect
is weakest at angles near 0 and 90, since the resolved shear stress due to the T-stress
vanishes.
As a practical matter, how large is the T-stress for a given geometry? To give
some insight, we use a finite crack of length 2a centered in an infinite solid with
remotely applied stress (see Figure 3). In this example, the applied stress intensity
factor K is a , and T is [3]. Combining and solving for T gives K / a .
W e m u s t c h o o s e s o m e K characteristic of dislocation nucleation, say,
2 u s/(1 ) , where is some dimensionless parameter of order unity or
moderately larger. Solving for the normalized T gives

(6)

Hence, we see that the T-stress effect may be most severe when the crack size is tens of
atomic spacings or less. More specifically, we have plotted the reduction in the
threshold for dislocation nucleation (as a percentage, from the uncorrected, or T=0,
result) versus crack size for several values of in Figure 4. We take b = h and = 0.3.
T-STRESS EFFECT ON DISLOCATION FORMATION 241

Figure 3. Finite crack of length 2a in an infinite solid, with remote tensile stress . The T-stress
for this particular geometry is ; thus, the threshold for dislocation nucleation is less than what
would be calculated using the classical K-field result.

For small values of as well as for small values of the crack size, the T-stress effect is
most significant. Clearly, one should proceed cautiously when applying continuum
concepts to cracks as small as, say, 10 atomic spacings; however, we note that the T-
stress effect can lead to appreciable reductions in Kc r i t for moderately longer cracks
when the slip plane inclination angle is less than about 45. This could lead to a
substantial source of error if trying to reconcile atomistic results for short cracks with
continuum-based predictions. If the crack is of some macroscopic dimension, then T
0 for this geometry, and the effect can be safely neglected. The results presented here
are consistent with the crack size effect on dislocation nucleation as noted by Beltz and
Fischer for mode III cracks [16] and Zhang and Li for mode I cracks [17].
We would like to dedicate this paper to Professor James R. Rice, with whom the
first author (GEB) had the great fortune to have as a mentor beginning ten years ago.
Jims high standards of scholarship, superb teaching, and supportive and modest
personality benefitted all of us whove had the chance to work with him.

4. References
1. Williams, M.L.: On the stress distribution at the base of a stationary crack, J. Applied Mechanics 24
(1957), 109-114.
2. Larsson, S.G., and Carlsson, A.J.: Influence of non-singular stress terms and specimen geometry on
small-scale yielding at crack tips in elastic-plastic materials, J. Mech. Phys. Solids 21 (1973), 263-277.
3. Rice, J.R.: Limitations to the small scale yielding approximation for crack tip plasticity, J. Mech. Phys.
Solids 22 (1974), 17-26.
4. Betegn, C., and Hancock, J.W.: Two-parameter characterization of elastic-plastic crack-tip fields, J.
Applied Mechanics 58 (1991), 104-110.
5. Fleck, N.: Finite element analysis of plasticity-induced crack closure under plane strain conditions,
Engineering Fracture Mech. 25 (1986), 441-449.
6. Cotterell, B., and Rice, J.R.: Slightly curved or kinked cracks, Int. J. Fracture 16 (1980), 155-169.
7. Fleck, N.A., Hutchinson, J.W., and Suo, Z.: Crack path selection in a brittle adhesive layer, Int. J.
Solids Structures 27 (1991), 1683- 1703.
242 G.E. BELTZ AND L.L. FISCHER

Figure 4. Reduction (from that of the T=0 result) of threshold for dislocation nucleation for a finite crack
of length 2a subject to remote tension (see Figure 3).

8. Langer, J.S., and Lobkovsky, A.E.: Critical examination of cohesive-zone models in the theory of
dynamic fracture, J. Mech. Phys. Solids 46 (1998), 1521-1556.
9. Peierls, R.E.: The size of a dislocation, Proc. Phys.Soc. 52 (1940), 34-37.
10. Nabarro, F.R.N.: Dislocations in a simple cubic lattice, Proc. Phys. Soc. 59 (1947), 256-272.
11. Rice, J.R.: Dislocation nucleation from a crack tip: an analysis based on the Peierls concept, J. Mech.
Phys. Solids 40 (1992), 239-271.
12. Frenkel, J.: Zur theorie der elastizittsgrenze und der Festigkeit Kristallinischer Krper, Zeitschrift
Phyzik 37 (1926), 572-609.
13. Beltz, G.E.: The Mechanics of Dislocation Nucleation at a Crack Tip, Ph.D. Thesis, Harvard
University, Cambridge, Massachusetts, 1992.
14. Beltz, G.E., and Rice, J.R.: Dislocation nucleation at metal-ceramic interfaces, Acta metall. mater. 40
(1992), S321-S331.
15. Rice, J.R., Beltz, G.E., and Sun, Y.: Peierls framework for dislocation nucleation at a crack tip, in A.S.
Argon (ed.), Topics in Fracture and Fatigue, Springer-Verlag, New York, 1992, pp. l-58.
16. Beltz, G.E., and Fischer, L.L.: Effect of finite crack length and blunting on dislocation nucleation in
mode III, Phil. Mag. A 79 (1999), 1367-1378.
17. Zhang, T.-Y., and Li, J.C.M.: Image forces and shielding effects of an edge dislocation near a finite
length crack, Acta metall. mater. 39 (1991), 2739-2744.
ELASTIC-PLASTIC CRACK GROWTH ALONG DUCTILE/DUCTILE
INTERFACES

W. J. DRUGAN
Department of Engineering Physics
University of WisconsinMadison
1500 Engineering Drive
Madison, WI 53706

Abstract: An analytical study is performed of the stress and deformation fields near the
tip of a crack that grows quasi-statically along an interface between two generally
dissimilar ductile materials. The materials are modeled as homogeneous, isotropic,
incompressible, elastic-ideally plastic Prandtl-Reuss-Mises, and the analysis is carried
out within a small-displacement-gradient formulation. The case of anti-plane shear
deformations is considered first. We derive near-tip solutions for the full range of the
ratio of the two materials yield stresses, and show that a near-tip family of solutions
exists for each set of material properties; the implication is that far-field loading and
geometrical conditions determine which specific near-tip solution governs in a
particular problem. As a by-product of this analysis, we derive a new solution family
for anti-plane shear crack growth in homogeneous material, one limiting member of
which is the familiar Chitaley and McClintock (1971) solution. We also analyze the
case of plane strain crack growth under applied tensile loading. Here, we account for
curvature of inter-sector boundaries, in an attempt to obtain a complete set of solutions.
When the material properties are identical, the solution family of Drugan and Chen
(1989) for homogeneous material crack growth, which has an undetermined parameter
in the near-tip field, is recovered. As the ratio of the two materials yield strengths,
deviates from unity, the near-tip solution structure is found to change, but the near-tip
fields are shown to continue to possess a free parameter for a substantial range of
Below this range, a second solution structure develops for which the near-tip free
parameter has a restricted range of freedom. Finally, a third near-tip solution structure
develops for sufficiently low for which there are no free parameters. The
implications of these results appear to be that as the plastic yield strength mismatch of
the two materials becomes larger, far-field loading and geometry have increasingly
weaker effects on the leading-order near-tip fields, until finally a mismatch level is
reached beyond which far-field conditions no longer affect the leading-order fields.
However, conclusions are complicated by the fact that the analysis also implies the
radius of validity of the leading-order fields to decrease continuously with increasing
yield strength mismatch (beyond a certain level), so that below some value, it will
243
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 243-274.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
244 W. J. DRUGAN

become necessary to retain more than one term to describe the physical near-tip fields.
Although not specifically explored here, our analysis also allows comparison of the
effects of changing elastic and plastic properties of the two materials on crack growth
propensity, so that perhaps this analysis could assist in the optimization of interfacial
fracture properties.

1. Introduction

Interfaces between two ductile materials arise in numerous situations of practical


importance. Composite materials is an obvious example; others include the interface
between base and weld metals, or base metal and the heat-affected zone, in a weld; clad
and coated materials; solder joints; and boundaries between different phases of the
same material (e.g., an austenite/martensite boundary). When the materials involved
possess at least some ductility, fracture of the interface if it is sufficiently strong
would be anticipated to occur not by an immediate catastrophic propagation of a crack,
but rather such would be preceded by a regime of slow, stable crack growth. A clear
understanding of the mechanics of this part of the failure process seems essential to the
overall characterization of ductile/ductile interface fracture.
We shall employ asymptotic (near-tip) analysis of the infinitesimal-displacement-
gradient governing equations to obtain analytical solutions for the stress and
deformation fields near a crack that propagates along the interface joining two
(generally different) ductile materials. Each material will be modeled as homogeneous,
isotropic, incompressible, elastic-ideally plastic Prandtl-Reuss-Mises. Two cases will
be analyzed: crack growth under anti-plane shear, and under plane strain with
nominally tensile applied loading. An interesting feature of the results is that, in both
cases, a family of solutions is shown to exist at each ratio of the materials yield
strength. For the anti-plane shear crack growth case, this is true for all values of the
shear strength ratio, whereas for plane strain crack growth, this is true only for a range
of yield strength ratios with one endpoint being the homogeneous material limit.
Previous analytical studies have treated crack growth along the interface between a
ductile material and a brittle one, the first being those of Drugan (1991) and Ponte
Castaeda and Mataga (1991). Recent studies of the fields near a stationary crack on a
ductile/ductile interface include those of Ganti and Parks (1997) and Sham, Li and
Hancock (1999). Of course, pioneering work on the elastic-plastic analysis of stress and
deformation fields near stationary and growing cracks in homogeneous materials is due
to Rice (1967, 1968, 1974, 1982, 1987, etc.), Rice and Tracey (1973) and Rice et al.
(1980) [with independent contributions from Slepyan (1974) and Gao (1980)], and the
analysis presented here, as with many other studies in diverse and numerous subjects in
solid mechanics, employs the elegant framework Jim Rice developed and builds on his
remarkable physical and mathematical insights.
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 245

2. Formulation
We will employ a small-displacement-gradient formulation to analyze quasi-static crack
growth along an interface joining incompressible, homogeneous, isotropic elastic-
ideally plastic materials modeled via an incremental (flow) theory. The crack geometry
to be analyzed in all cases is illustrated in Figure 1. A Cartesian coordinate system is
fixed in the body such that x1 points in the crack growth direction and x 3 is parallel to
the (locally straight) crack front. A polar (r, ) coordinate system is centered at the
crack tip and moves with it during crack growth. Angle i s m e a s u r e d
counterclockwise from the x1 axis, and the crack growth speed is The crack is
shown growing along the interface between two generally different elastic-ideally
plastic materials, having yield stresses in pure shear ki and elastic shear moduli Gi .

Figure 1. Cartesian x 1 , x 2 coordinate system is fixed in the body; polar r, system is centered at the
crack tip and moves with it through the material during crack growth.

The governing equations within each homogeneous material are as follows.


Equilibrium requires the stress tensor to satisfy

(1)

where is the usual vectorial differentiation operator so that denotes divergence,


b is the body force vector per unit volume and a superscript T denotes tensor
transpose. The rate of deformation tensor D is related to the material velocity vector v
as

(2)
246 W. J. DRUGAN

Elastic-plastic response in each material is assumed to be described` by the Prandtl-


Reuss flow rule

(3)

where e and p denote elastic and plastic parts, respectively, G is the elastic shear
modulus, s = I trace( )/3 is the deviatoric stress tensor with I denoting the
second-rank identity tensor; 0 is an undetermined parameter and a superposed dot
denotes time rate at a fixed material point. Material is at yield when the Huber-Mises
condition is met:

(4)

where k is the (constant) yield stress in shear. Thus, in regions of ongoing plastic
deformation, (3) and (4) must be satisfied with > 0; in material experiencing
instantaneously elastic response, whether or not it has previously deformed plastically,
(3) governs with 0 and (4) applies with replacing = .
Rice (1982) argued, and Drugan (1985) proved, that these governing equations
become analytically tractable in the limit as r 0, even for general unsteady crack
growth. Specifically, if b is assumed bounded everywhere and mild, physically
sensible assumptions are made concerning the existence of certain quantities involving
stress derivatives as r 0, Rice (1982) and Drugan (1985) proved that (1) and (3)
simplify to the following asymptotic forms, specialized here to the present problems in
which all field quantities are assumed independent of x3 :

(5a)

(5b)

(5c)
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 247

(6)

respectively, where R is an undetermined parameter having length dimensions.


As will be reviewed explicitly in subsequent sections, in the cases of antiplane
strain and plane strain, the asymptotic (r 0) form of this governing equation set was
shown by Rice (1982) to admit three types of solution i.e., three types of near tip
angular sector, in a given material. The construction of complete leading-order
solutions for growing crack stress and deformation fields thus involves assembling
sectors of these permissible types in such a manner that plastic deformation is
everywhere nonnegative, that the yield condition is not violated in elastically deforming
sectors, and that the appropriate inter-sector and boundary conditions are satisfied.
Since we shall consider growing cracks, intersector boundaries (other than one
lying on = 0) propagate quasi-statically through the material. Drugan and Rice
(1984) analyzed restrictions on such quasi-statically propagating surfaces for a general
class of elastic-plastic materials, inclusive of the model employed here. Recently,
Drugan (1998a) provided rigorous confirmation of certain ad hoc assumptions they
employed, and generalized their results to an even broader class of materials.
Specialized to the present incompressible material model and the deformation modes
considered here, these analyses show that the following important conditions must be
satisfied across propagating intersector boundaries: (i) all components of the stress
tensor must be continuous; (ii) only the material velocity components parallel to the
moving surface, and their associated plastic shear strain components, may experience
jumps, but only if all of the following conditions are met:

(7a)

(7b)

(7c)

Here, a superscript indicates that the (continuous) stress component is to be


evaluated on the potential discontinuity surface; subscripts n and t refer to normal (in
the propagation direction) and tangential (in the x1 , x 2 plane) directions to the moving
surface whose normal velocity is c, and subscript stands for t and 3; [ ] denotes
the jump in a quantity across the surface (value just ahead minus value just behind);
and [ ] 0.
248 W. J. DRUGAN

In addition to the above inter-sector continuity conditions, we shall require the


minimum necessary conditions across the line ahead of the crack ( = 0), so as to
explore all possible near-tip solutions: namely, continuity of the traction and
displacement vectors will be enforced. (Note that continuity of displacement across
= 0 requires continuity of velocity across that line.) Finally, we shall impose the
condition of traction-free crack faces.

3. Anti-Plane Strain

3.1 GENERAL SOLUTION SECTOR TYPES AND CONTINUITY CONDITIONS

We shall first provide a detailed perhaps exhaustive study of the anti-plane strain
case, both because of its relative mathematical simplicity but also because observations
and trends evident here will lend great insight into the more complicated plane strain
case. In anti-plane strain, the only nonvanishing velocity component is v3 = v 3 (r, ) .
From this, (1) (3) show that the only nonvanishing components of and D are the
13 (=31) and 23 (=32) ones, which must also be independent of x3 . Thus, the yield
condition (4) reduces to

(8)

As reviewed lucidly by Rice (1982), the leading-order in r as r 0 forms of the


governing equations, namely (2), (5c), (6) and (8), admit three types of solution i.e.,
three types of near tip angular sector. These are summarized below, with the sign of
3 in the centered fan plastic sector chosen positive, as it will be in all our solutions.

(i) Centered Fan Plastically Deforming Sector

(9)

(10)

(11)
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 249

(ii) Constant Stress Plastically Deforming Sector

(12)

(13)

(14)

(iii) Elastically Deforming Sector

(15)

(16)

In the preceding equations, A, B, C, Q 3 and V are undetermined dimensionless


constants except that Q 3 must be chosen such that (12) satisfies (8); subscript has
range 1, 2; G is the elastic shear modulus; ( , t) is an undetermined function of
integration; and t is time or some monotonically increasing parameter. Equations
(13), (14) were derived by employing (6) and assuming that the velocity field in a
constant stress plastically deforming sector is logarithmically singular in r, as it must
be in the other sector types. This derivation is similar to its plane strain counterpart,
which is given in the Appendix.
Acceptable solutions to the near-tip growing crack fields must be assembled from
the allowable solution sector types given above within each material, with the
appropriate continuity/jump conditions enforced across inter-sector boundaries and the
appropriate boundary conditions applied. For all intersector boundaries excluding
= 0, the general conditions described in Section 2 specialize in the anti-plane strain
case to the following:

[ 3 ] = 0 (17)

(18)

Across = 0, we require continuity of 23 (= 3 ) and of v 3 , and on = we


require 23 = 0.
250 W. J. DRUGAN

3.2 SOLUTIONS FOR DUCTILE/DUCTILE INTERFACE CRACK GROWTH

There appear to be two general near-tip solution configurations for anti-plane shear
crack growth along a ductile/ductile interface between two elastic-ideally plastic
materials that in general have different elastic and plastic properties. These are
illustrated in Figure 2.

Figure 2. Near-tip solution configurations for anti-plane shear ductile/ductile interface crack growth.
Material 1 lies above the crack line and has shear strength k 1, while Material 2 lies below and has
shear strength k 2 . Sectors A, B, G, H are constant stress plastic; C, D are centered fan plastic;
and E, F are instantaneously elastic.
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 251

In both configurations of Figure 2, the boundaries 1 and 2 are stress characteristics,


in Figure 2(a) because of full stress continuity with centered fans, and in Figure 2(b)
because these boundaries are sites of velocity jumps. This means that the stress states in
the constant stress plastic sectors A and B have the following representations (the
superscript denotes the appropriate sector):

(19)

(20)

Thus, in both configurations, traction continuity across = 0 requires

(21)

This shows that 1 and 2 can only be equal for crack growth in homogeneous
material, so that at least one of the constant stress sectors A or B is necessary in
general for interfacial crack growth along yield-strength-mismatched interfaces.
Incidentally, it is easy to see that a sub-yield elastic sector cannot border = 0 in any
solution: nonviolation of the yield condition would require B = 0 in (15), rendering the
stresses constant; then, full stress continuity with a necessarily at-yield trailing sector
would result in the stress field in the hypothesized elastic sector being at yield.
Proceeding with the analysis of the Figure 2(a) configuration: full stress continuity
is required across 1 and 2, which is already satisfied by (19) and (20). From (18),
because these boundaries are characteristics, i.e. have r3 = 0, they are permitted to
have jumps in v 3 , which indeed are necessary to avoid infinite velocities in Sectors A
and B as these boundaries are approached. That is, (13) shows that as approaches
1 in Sector A or 2 in Sector B, v 3 becomes infinite unless V A = V B = 0. This
choice being made, it is clear that leading-order velocity continuity is satisfied across
= 0 (since velocities are zero at leading order in Sectors A, B). Observe that the v 3
jumps across 1 and 2 produce positive plastic work.
Next, we enforce full stress and velocity continuity across the remaining intersector
boundaries. As noted earlier, stress continuity is required; velocity continuity is
required across 5 and 6 because the condition r3 = 0 will not be met on these
boundaries. Although this condition is met on 3 and 4 , one can prove that the
requirements of positive plastic work produced by a hypothesized velocity jump,
together with the stress state in the elastic sector not violating yield, rule out a velocity
jump across a propagating centered fan/elastic sector boundary [the plane strain version
of this proof, due to Rice, is given in Appendix A of Drugan and Chen (1989)]. These
continuity conditions lead to equations for the boundary angles and the constants
appearing in (12) (16) for Sectors E H that do not involve the elastic shear moduli
nor the yield strengths. The equations governing the boundary angles are:
252 W. J. DRUGAN

(22a, b)

with identical equations for angles 4 a n d 6 , except that the last +1 in (22a) is
replaced by 1. The unique physically meaningful solutions to these equations are:

3 = 4 = 19.711224 , 5 = 6 = 179.63341; (23)

Figure 2(a) has been labeled with these unique angles (although it is not drawn with
these angles for purposes of illustration). The other constants appearing in (12) (16)
for Sectors E H are, where a superscript again denotes the sector of applicability of
the constant:

(24)

Thus, for the range of applicability of this solution family (about which, more shortly),
all boundary angles except 1 a n d 2 are fixed and independent of the far-field
loading and mismatch in material properties, as are the angular variations of the stress
and deformation fields in all sectors except Sectors A and B. Different material
properties change the amplitudes of the stress and deformation fields, by changing the
amplitudes of k and G that appear in (9) (16).
To determine the range of material property mismatches for which this solution
family can apply, we note that (21) and (23) must both be satisfied. The first interesting
conclusion, then, is that, since the materials elastic shear moduli do not enter (21), this
solution family appears capable of providing solutions for any elastic mismatch (at least
the asymptotics suggests this). However, simultaneous satisfaction of (21) and (23)
does limit the ratio of yield strengths for which this solution type applies, since (23)
shows that 19.711224 2 0 1 19.711224 . This with (21) shows that the
solution configuration of Figure 2(a) can provide interface crack growth solutions for
the yield strength ratio range:

(25)
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 253

One observes that for specific values of k2 /k1 inside this range, 2 is strictly specified
in terms of 1 , but there is some freedom of choice of one of these. The specific
solution must be determined from far-field conditions of the particular problem under
consideration. When k 2 /k1 = 1.0622426 (0.94140449), this corresponds to 1 = 0,
2 = 19.711224 ( 1 = 19.711224 , 2 = 0): in these cases, say the first one for
definiteness, Sector A has disappeared, as has Sector D. This limit case is illustrated
in Figure 3.

Figure 3. Near-tip solution configuration for k2 /k1 = 1.0622426.

A s k 2 /k 1 increases above 1.0622426, the configuration shown in Figure 3 will still


provide near-tip solutions; interestingly, the fields throughout Material 1 remain
unchanged in their angular distribution, but in Material 2, angles 4 and 6 , and hence
the fields in Sectors B, F and H, change as k 2 /k1 increases. The equations governing
the values of these angles and the parameters in the stress and velocity fields in these
sectors are obtained by writing (21) with 1 = 0 and 2 = 4 , and enforcing full stress
continuity across 4 and 6 , and velocity continuity across 6 . Note that 4 is now
the site of a velocity discontinuity, which produces positive plastic work and
corresponds to an elastic sector stress field that does not violate yield. The results are:

(26)

(27)
254 W. J. DRUGAN

(28)

Solutions to (26) (28) are listed in Table 1 for the full allowable range of
1.0622426 k 2 /k1 (reported as 0 k1/ k2 0.94140449).

k 1/k2 4 6 AF BF CF

0.9414 19.71 179.633 1.057 0.3373 0.7038

0.9 25.84 179.655 1.051 0.3353 0.7143

0.8 36.87 179.766 1.006 0.3208 0.7639

0.7 45.57 179.868 0.9375 0.2986 0.8147

0.6 53.13 179.940 0.8514 0.2711 0.8613

0.5 60 179.980 0.7500 0.2388 0.9007

0.4 66.42 179.996 0.6339 0.2018 0.9342

0.3 72.54 179.9997 0.5025 0.1600 0.9615

0.2 78.46 1 8 0 0.3546 0.1129 0.9821

0.1 84.26 1 80 0.1880 0.0598 0.9953

0.0 90 180 0 0 1

Table 1. Values of the parameters for the solution configuration of Figure 3, for 0 k1/k 2 0.94140449.

Interestingly, solutions for this range of k l /k2 can also be found with the near-tip
configuration of Figure 2(b). The analysis of this configuration proceeds as follows:
As explained earlier, the boundaries 1 and 2 are stress characteristics because they
are the sites of velocity jumps; the stress states in Sectors A and B are therefore given
by (19) and (20), and for the reasons explained earlier, V A = V B = 0. Thus, the
velocities are zero in these sectors, and hence velocity continuity is satisfied across
= 0. Full stress continuity is enforced across all intersector boundaries (except = 0,
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 255

which already has traction continuity satisfied), and velocity continuity is enforced
across 3 and 4 . The resulting equations are:

(29)

(30)

(31)

(32)

(33)

Thus, for each value of the ratio k 1/k 2 0.94140449, a range of solutions having this
configuration can be found. The limitations are that

1 19.71 and 2 19.71; (34)

this is because otherwise the yield condition would be violated in the elastic sectors.
For example, for the case k1 /k 2 = 0.8, (29) and (34) show that there will be a range of
solutions with 19.71 1 9 0 and 41.14 2 90, with the other angles and
constants being given by (30) (33). A specific set of angles 1, 2, satisfying (29) and
(34), apparently is determined by the far-field boundary and loading conditions of the
specific problem under consideration.

3.3 NEW SOLUTION FAMILY FOR CRACK GROWTH IN HOMOGENEOUS


MATERIAL

We now specialize the preceding analysis to exhibit a new solution family for anti-plane
strain crack growth in homogeneous materials. To parameterize the solutions within this
family, it will be convenient to introduce a near-tip mixity parameter M as:
256 W. J. DRUGAN

(35)

The new solution family has solutions for all mixities in the range 0 < M 1. The
solution family has the two configurations illustrated in Figure 2, except that now, of
course, the yield strengths are the same in the upper and lower materials. This fact
reduces (29) to the requirement that

1 = 2 . (36)

This requirement applies to both configurations of Figure 2.


The remaining features of the solutions are very similar to those just analyzed for
crack growth along an interface involving two different materials. Specifically: from
(19), (20), (35), (36) and the fact that 3 = 4 = 19.71, we deduce that the solution
configuration illustrated in Figure 2(a) applies for mixities in the range 1 M 0.7810.
Notice that M = 1 is the Chitaley-McClintock (1971) solution, which has fully
continuous stresses and velocities. As M decreases from this value to 0.7810, 1 (and
2) increase from zero to 19.71 . As discussed earlier, this only changes the solution
fields in the range 2 1 , but the solution now has (admissible) velocity jumps
across 1 and 2 (and zero leading-order velocities within Sectors A and B), and a
radial shear stress jump across = 0. The stress and velocity fields for solutions of this
type are given by (9) (16) and (19), (20) together with the angles and parameter values
given by (23), (24) and the combination of (35) and (19). For mixities in the range
0.7810 M > 0, the solution adopts the configuration of Figure 2(b), but satisfying (36).
As above, the angle 1 (= 2 ) is determined from (19) and (35) for a given mixity,
and the remaining angles and parameter values are given by (30) (33), which are
simpler now since the fields are antisymmetric about = 0. For the range of mixities
just given, 1 sweeps through the range 19.71 1 < 90 . Which member of this
solution family applies (i.e., which mixity) is determined by far-field loading and
geometry. The small-scale yielding numerical finite element solutions of Dean and
Hutchinson (1980) and Freund and Douglas (1982), together with the additional
analysis of Drugan (1998b), seem to imply the near-tip presence of the M = 1 solution in
the small-scale yielding case, but it seems that e.g. crack growth under general yielding
conditions, or even under contained yielding conditions but with the second term in the
Williams expansion for the surrounding elastic fields nonzero, may well produce others
of the solution members derived here. Certainly, all results presented here are
completely admissible solutions to the asymptotic (near-tip) governing equations.
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 257

4. Plane Strain

4.1 GENERAL SOLUTION SECTOR TYPES AND CONTINUITY CONDITIONS

In plane strain, the nonzero components of material velocity are v 1 = v l ( r, ),


v2 = v2 (r, ). From these, (1) (3) show that the only nonvanishing components of s
and D are the 11, 12 (=21) and 22 ones [and 33 = ( 11 + 22 )/2], which must also be
independent of x 3. Thus, the yield condition (4) reduces to

(37)

As derived elegantly by Rice (1982), the leading-order in r as r 0 forms of the


governing equations (2), (5a), (5b), (6) and (37), admit three types of solution i.e.,
three types of near tip angular sector. These are summarized below.

(i) Centered Fan Plastically Deforming Sector

(38)

(39)

(40)

(41)

(ii) Constant Stress Plastically Deforming Sector

(42)

(43)

(44)
258 W. J. DRUGAN

(45)

(46)

(iii) Elastically Deforming Sector

(47)

(48)

(49)

(50)

(51)

(52)

In the preceding equations, A , B , C, C , P a n d V are undetermined


dimensionless constants except that P must be chosen such that (42) satisfies (37);
subscripts , , have range 1,2 and obey the summation convention; G is again the
elastic shear modulus; f (, t ) is an undetermined function of integration; e is the
natural logarithm base and is the Kronecker delta.
Equations (43)(46) were derived by assuming that the velocity field in a constant
stress plastically deforming sector is logarithmically singular in r, as it must be in the
other sector types. A new, direct derivation of these is provided in the Appendix.
Acceptable solutions to the near-tip growing crack fields must be assembled from
the allowable solution sector types given above in each material, with the appropriate
continuity/jump conditions enforced across inter-sector boundaries (including across the
material interface) and the boundary conditions of zero tractions on the crack faces. For
all intersector boundaries excluding = 0, the general conditions described in Section 2
specialize in the plane strain case to the following:
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 259

[ ] = [ 33 ] = 0 (53)

[ v ] = 0 (54)

(55)

4.2 SOLUTIONS FOR DUCTILE/DUCTILE INTERFACE CRACK GROWTH

Knowing that for plane strain crack growth in homogeneous materials there are two
solution types Mode I or tensile fields, and Mode II or shear fields we anticipate two
solution types for the interface crack growth problem also. We here focus on tensile-
type solutions, anticipating that they will be the more important physically. By tensile-
type solutions, we refer to solutions that reduce to the Mode I solutions when the
properties of the two materials become identical; when these materials are different,
these tensile-type solutions will of course not be symmetric or pure Mode I, as the
ensuing analysis shows.

Figure 4. Solution configuration for plane strain ductile/ductile interface crack growth.
Material 1 lies above the crack line, Material 2 below.

We begin with the analysis of the configuration shown in Figure 4 for the tension-
type case. This means that we choose r positive in Sector C and negative in Sector
D. Then, enforcing full stress continuity across 1 and 2 , as required by (53), the
stress fields in Sectors A and B are, employing (38) and (42) in each material (again,
superscripts denote the applicable sector):
260 W. J. DRUGAN

(56)

(57)

Enforcement of traction continuity across = 0 thus requires

(58)

(59)

The forms (56), (57) also permit specification of the velocity fields in Sectors A
and B. First, for Sector A, (46) become

S = cot 2 1 , b l =cot l , b 2 = tan l ; (60)

using these, (43) (45) are, for Sector A:

(61)

(62)

(63)

These have exactly the same forms in Sector B, except that the subscripts of k, G, are
changed from 1 to 2, and superscripts A are changed to B. Leading-order velocity
continuity (i.e., analyzing only the logarithmically singular terms) across = 0 thus
requires (when 1 0 and 2 0):
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 261

(64)

(65)

Now, in their analysis of tensile crack growth in homogeneous materials, Drugan


and Chen (1989) derived a family of leading-order solutions for the growing crack stress
and deformation fields by accounting for curvature of the boundaries that in Figure 4
have the angular locations 1 and 2 as r 0. We shall perform an equivalently
general analysis here. This means that the boundaries just mentioned will be taken to
have the asymptotic forms, respectively, where m is an initially unspecified parameter:

(66a, b)

When the boundaries have these forms, there is an additional In(R/r) contribution to
the Sector A velocity fields as r 0 along the Sector A/C boundary from the
terms in (61), (62), and similarly for Sector B; this introduces an
additional (free) parameter m into the leading-order solution. We have chosen the
same power m for the boundaries in the upper and lower materials because, as Drugan
and Chen (1989) showed, boundary forms of type (66) lead to second-order stress field
corrections proportional to (r/R) m , and we require traction continuity across = 0.
We shall seek growing crack solutions with full stress and velocity continuity
across all intersector boundaries (except for possible rr jumps across = 0); the
homogeneous material solutions found by Drugan and Chen (1989) have this property
(except for the m limiting case, as will also be true here). Thus, employing (66)
in (61), (62) and (39), (40), full leading-order velocity continuity across the Sector A/C
and B/D boundaries requires:

(67)

(68)

(69)

(70)
262 W. J. DRUGAN

We also enforce full stress and velocity continuity across all other remaining intersector
boundaries; jumps in v r across 3 and/or 4 seem possible initially but are ruled out
by the proof in Appendix A of Drugan and Chen (1989). The equation system resulting
from all these additional continuity conditions can be summarized as follows. In
Material 1:

(71)

(72)

(73)

(74)

(75)

(76)

(77)

(78)

The equations and parameters in Material 2 ( < 0) are given by equations identical in
form to (71)(78), with the following changes: superscripts (D, F, H) replace (C, E, G);
angles (2 , 4 , 6 ) replace ( 1, 3 , 5 ) (and of course negative solution values are
sought); and the signs are changed of: the right sides of (71), the last sin term in (72),
the last cos term in (73), and 2 3 becomes 2 4 in (74). Thus the total equation
system to be solved is (58), (59), (64), (65), (67)(70), (71)(78), and the Material 2
versions of (71)(78) just explained. There is one more unknown than the total number
of equations, so that at least throughout the range for which the configuration of
Figure 4 prevails, a one-parameter family of solutions is expected for each set of
material properties. This is indeed found to be the case. Following Drugan and Chen
(1989), we regard m as the undetermined parameter.
This system is solved as follows: First, one solves (67)(70) for
in terms of V C, V D; substitution of these results into (64), (65) then permits solution of
V C, V D in terms of angles and m only. The results are:
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 263

(79)

(80)

(81)

(82)

where = k l / k2 and = G 1 / G 2 . Using (79)(82) in (71), (74) and their Material 2


counterparts, the six equations governing the six angles in Figure 4 are (58), (59), (72),
(73) and the Material 2 counterparts of the latter two.
Our approach was to begin with the two ductile materials being identical the case
for which Drugan and Chens (1989) homogeneous material solutions apply. Solutions
for different ductile materials can then be obtained by slowly changing the values of
yield stress and/or elastic moduli of the two materials, while tracking the solutions
whose structure initially resembles that of Drugan and Chen (1989). We will
summarize a few key results here to give a flavor of what the solutions show, focusing
on the effect of changing the yield stress ratio between the two materials (i.e., we have
chosen G 1 = G 2 in the solutions).
For the purpose of obtaining specific results, we begin by choosing the value of
Drugan and Chens (1989) near-tip characterizing parameter m = 1.24, the value Liu
and Drugan (1993) showed it takes for homogeneous crack growth when small-scale
yielding conditions prevail (i.e., the plastic zone size is small compared to the crack
length and the distance to specimen boundaries). Figure 5 shows the leading-order (in
distance, r, from the crack tip) near-tip solution structure for this m-value when the two
materials are identical, as determined by Drugan and Chen (1989). The two families of
orthogonal lines in plastic regions are stress characteristics; a material element with
sides parallel and perpendicular to such a line has a stress state consisting of shear stress
equal to yield stress in pure shear, plus a hydrostatic stress. Sectors A, B, G and H are
constant stress plastic sectors, in which the Cartesian components of stress do not
vary with position; Sectors C and D are centered fan plastic sectors, in which the
stress state consists of the polar shear component of stress equaling the materials yield
stress in shear, plus a hydrostatic stress that varies linearly with angle; and Sectors E
and F are instantaneously elastic, although containing material which has previously
deformed plastically. The values of the inter-sector boundary angles, 1 6, are given
by the first row in Table 2, and Figure 5 is drawn with these angles.
264 W. J. DRUGAN

Figure 5. The near-tip solution configuration for homogeneous material,


i.e., =1 with m = 1.24 (boundary angles drawn to scale).

Now recalling that = k l / k 2, i.e., the ratio of the upper material (x2 > 0) to lower
material (x 2 < 0) yield stress, we found first that as decreases from 1 to about 0.657,
the near-tip solution structure resembles that of Figure 5 except that the inter-sector
boundary angles change smoothly with decreasing . Table 2 shows how these angles
change with , and Figure 6 illustrates the near-tip configuration for = 0.7.
Observe that the region of strongest plastic strain ing in each material, the centered fan
sector, has drastically enlarged in the upper (lower yield stress) material, and
dramatically condensed in the lower (greater yield stress) material. Also, the sector of
purely elastic response has greatly enlarged in this lower material. For any of the values
of given in Table 2, the solutions for the angles listed there can be employed to
easily determine the values of all the other parameters appearing in the near-tip stress
and deformation fields, via use of (79)(82), (71), (74)(78) and the Material 2
cou nterparts of these last six equations. Once these parameters have been calculated,
the actual stress and velocity fields are given by (38)(52), choosing the appropriate
ones of these for each particular sector, and ensuring that the material properties
correspond to whether the fields are in Material 1 or 2.
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 265

1 2 3 4 5 6

1 45 45 118.927 118.927 156.860 156.860

0.9 36.9723 52.2066 121.053 115.802 154.869 159.451

0.8 26.3465 59.5019 122.778 109.666 153.098 163.603

0.7 9.71432 65.6553 124.288 87.9485 151.422 172.256

0.657166 4.31851 65.2600 124.764 65.2600 150.867 176.434

Table 2. The near-tip inter-sector boundary angles (measured anticlockwise from x1, in degrees)
as functions of , for the near-tip configuration of Figures 5 and 6 (m = 1.24).

Figure 6. The near-tip solution configuration for = 0.7 with m = 1.24


(boundary angles drawn to scale).
266 W. J. DRUGAN

In addition to analyzing how the near-tip solution configuration alters as the yield
stress ratio changes for the small-scale yielding value of m = 1.24, we examined
whether the range of m for which solutions could be found at each value was
altered by changing The importance of this is as follows: m is the only parameter
undetermined by the asymptotic analysis in the leading-order (in r) asymptotic solution,
and thus the near-tip fields experience differences in far-field loading, geometry and
yielding extent only through changes in m. Furthermore, the next term in r in the
asymptotic stress field derived by Drugan and Chen (1989) is proportional to (r / R) m ,
where R is a parameter with length dimensions that also depends on far-field
conditions [look ahead to (92)]. Thus m also regulates how rapidly the leading-order
stress fields alter with distance from the crack tip. For Mode I crack growth in
homogeneous material, Drugan and Chen (1989) showed that near-tip solutions can be
found for all m -values in the range 0 < m .
Our analysis of the present problem led to a fascinating result: For 1 0.7,
we find that near-tip solutions are possible for all m-values in the range of acceptable
solutions in homogeneous materials, i.e., 0 < m . However, as becomes slightly
less than about 0.7, the range of m-values for which near-tip solutions can be found is
dramatically narrowed, since the maximum admissible m-value is drastically decreased!
This is illustrated in Table 3, which displays the maximum m for which an acceptable
near-tip solution can be found, as a function of

m max

0.7 to 1
0.69354 27,5676.0
0.69 14.2562
0.68 4.39663
0.67 2.99652
0.65 2.18306
0.60 1.32282
0.591622 1.24
Table 3. Maximum m-value for which near-tip solutions can be found
(continues in Table 5).

Returning now to our solutions for m = 1.24, observe from Table 2 that when
reaches the value 0.657166, 2 a n d 4 have coincided, meaning that Sector D, the
centered fan sector in the lower material, has disappeared. The near-tip solution
configuration has thus become that illustrated in Figure 7. There is still one more
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 267

parameter than the number of equations constraining them for this configuration, since
now the restrictive form of the velocity field in Sector D is no longer imposed.

Figure 7. Near-tip configuration for = 0.657166, drawn to scale (m = 1.24).

Specifically, the Material 2 versions of (71) no longer apply, nor do (69) (70),
which were the requirements of velocity continuity across 2 . The new equations of
velocity continuity across 2 are, employing (51), (52), the Sector B versions of (61),
(62), and (66b):

(83)

Equations (81), which resulted from (67)(68), still apply; using these in (64)(65) and
solving the resulting equations for gives:

(84a)

(84b)
268 W. J. DRUGAN

Since 2 is a characteristic, its stress state can still be expressed as if it were a centered
fan of zero angular extent, so that full stress continuity across 2 still results in the
stress field in Sector B being expressible as (57). Therefore, (58)(59) still apply. The
full equation set is thus (58), (59), (71)(78), the Material 2 versions of (72)(78) except
that now 2 replaces 4 , (81), (83) and (84). This is reduced to five equations for the
five undetermined angles as follows: use (84) in (83) to express the in terms of
V C . Use these and (71) in (74) and its Material 2 counterpart to express CC and C D in
terms of V C. Substitution of these into (59) gives an equation that can be solved for VC
involving only boundary angles; the result is:

(85)

Then, employing (85) with (71), (83) and (84), allow (72), (73) and their Material 2
versions to be expressed purely in terms of the boundary angles, as is (58); these are the
five equations to be solved. The solution of these equations shows that the
configuration of Figure 7 persists until attains the value 0.591622, at which value
1 has reached the value 0. Table 4 gives the near-tip sector boundary angles for
corresponding to solutions with the configuration of Figure 7.

1 2 3 5 6

0.657166 4.31851 65.2600 124.764 150.867 176.434

0.6 0.483300 63.4319 124.749 150.885 178.754

0.591622 0 63.1361 124.749 150.885 178.954

Table 4. Near-tip inter-sector boundary angles (measured anticlockwise from x1 , in degrees)


as functions of for the near-tip configuration of Figure 7 (m = 1.24).
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 269

For 0.591622, Sector A, the leading constant stress sector in the upper
material, has disappeared, so that the near-tip configuration is that shown in Figure 8.

Figure 8. Solution configuration for 0.591622, drawn to scale for the case of equality.

The analysis of this configuration proceeds as follows. Equations (57) are still
valid, and (58), (59) apply with 1 = 0; these become:

(86)

(87)

Equations (64), (65) no longer apply; the new conditions of velocity continuity across
= 0 are, using (39), (40) and the Sector B version of (61), (62):

(88)
270 W. J. DRUGAN

(89)

Solving (88), (89) for gives

(90a, b)

The full equation set is thus (86), (87), (90), (71)(78), the Material 2 versions of
(72)(78) except that now 2 replaces 4, and (83). This is reduced to equations for
the undetermined angles as follows: use (90) in (83) to express the in terms of V C .
Use these and (71), in (74) and its Material 2 counterpart, to express C C and C D in
terms of V C . Substitution of these into (87) gives an equation that can be solved for VC
involving only boundary angles; the result is:

(91)

Then, employing (91) with (71), (83) and (90), allow (72), (73) and their Material 2
versions to be expressed purely in terms of the boundary angles, as is (86); these are the
five equations to be solved. Observe, however, that there are only four remaining
angles to be determined in the present solution configuration (Figure 8)! This means
that the near-tip parameter m is no longer a free parameter for < 0.591622: only
one near-tip solution can be found for each in this range. The solutions of the just-
described five-equation set are reported in Table 5.
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 271

m 2 3 5 6

0.591622 1.24 63.1361 124.749 150.885 178.954

0.55 0.939418 61.6835 125.179 150.372 179.603

0.5 0.713333 60 125.529 149.948 179.910

0.45 0.558467 58.3718 125.776 149.643 179.988

0.4 0.442601 56.7891 125.960 149.413 179.999

0.35 0.351256 55.2437 126.101 149.236 179.999993

0.3 0.276698 53.7288 126.211 149.097 179.999999995

0.25 0.214174 52.2388 126.297 148.987 180 + 2.1 1014

0.2 0.160548 50.7685 126.365 148.901 180 + 9.7 1025

0.15 0.113652 49.3135 126.418 148.834 180 + 9.3 1048

0.1 0.071932 47.8696 126.459 148.781 180 + 8.1 10116

0.05 0.034242 46.433 126.490 148.741 180 + 3.1 10497

Table 5. Near-tip inter-sector boundary angles and the unique m-value for which a solution exists, as
functions of 0.591622, for the near-tip configuration of Figure 8.

One extremely important implication of these results is as follows: Drugan and


Chen (1989) showed that in terms of a polar coordinate system centered at the moving
crack tip, the stress field in a centered fan plastic sector has the structure (and this is
anticipated to be the case in the other near-tip sectors):

(92)

where s (0)( ; m), s (1)( ; m ), , are specified by the asymptotic analysis, except for the
value of the characterizing parameter m (and additional constants in the higher-order
272 W. J. DRUGAN

terms). The results summarized in Table 5 thus show that for sufficiently large , the
near-tip fields may be accurately described simply by the leading-order stress field [i.e.,
by just one term in (92)]. However, for values below a certain level, m has become
sufficiently small that a leading-order analysis will certainly not suffice, so that accurate
representation of physical near-tip stress fields will require retention of more than one
term in (92). The value below which more than one term must be retained depends
on the size scale of the microstructure of the material being modeled (since the radius of
validity of the near-tip fields must be larger than this microstructural size, and the radius
of validity of the leading-order term depends on m, and hence on ). The same
conclusion applies for the near-tip deformation fields.

Acknowledgements

Support of the National Science Foundation, Mechanics and Materials Program, under
Grant CMS-9800157 is gratefully acknowledged.

References

Chitaley, A. D. and McClintock, F. A. (1971), Local Criteria for Ductile Fracture, Journal of the Mechanics
and Physics of Solids, Vol. 19, pp. 147- 163.
Dean, R. H. and Hutchinson, J. W. (1980), Quasi-Static Steady Crack Growth in Small Scale Yielding,
Fracture Mechanics, ASTM-STP 700, pp. 383-405.
Drugan, W. J. (1985), On the Asymptotic Continuum Analysis of Quasi-Static Elastic-Plastic Crack Growth
and Related Problems, Journal of Applied Mechanics, Vol. 52, pp. 60l-605.
Drugan, W. J. (1991), Near-Tip Fields for Quasi-Static Crack Growth Along a Ductile-Brittle Interface,
Journal of Applied Mechanics, Vol. 58, pp. 111-119.
Drugan, W. J. (1998a), Thermodynamic Equivalence of Steady-State Shocks and Smooth Waves in General
Media; Applications to Elastic-Plastic Shocks and Dynamic Fracture, Journal of the Mechanics and
Physics of Solids, Vol. 46, pp. 313-336.
Drugan, W. J. (1998b), Limitations to Leading-Order Asymptotic Solutions for Elastic-Plastic Crack
Growth, Journal of the Mechanics and Physics of Solids, Vol. 46, pp. 2361-2386.
Drugan, W. J. and Chen, Xing-Yu (1989), Plane Strain Elastic-Ideally Plastic Crack Fields for Mode I
Quasistatic Growth at Large-Scale Yielding I. A New Family of Analytical Solutions, Journal of the
Mechanics and Physics of Solids, Vol. 37, pp. l-26.
Drugan, W. J. and Rice, J. R. (1984), Restrictions on Quasi-Statically Moving Surfaces of Strong
Discontinuity in Elastic-Plastic Solids, Drucker Anniversary Volume: Mechanics of Material Behavior,
edited by G. J. Dvorak and R. T. Shield, Elsevier Scientific Publishing Company, Amsterdam, pp. 59-
73.
Freund, L. B. and Douglas, A. S. (1982), The Influence of Inertia on Elastic-Plastic Antiplane-Shear Crack
Growth, Journal of the Mechanics and Physics of Solids, Vol. 30, pp. 59-74.
Ganti, S. and Parks, D. M. (1997), Elastic-Plastic Fracture Mechanics of Strength-Mismatched Interface
Cracks, in Mahidhara, R. K. et al. (Eds.), Recent Advances in Fracture, The Minerals, Metals and
Materials Society, pp. 13-25.
Gao, Y. C. (1980), Elastic-Plastic Field at the Tip of a Crack Growing Steadily in Perfectly Plastic Medium
(in Chinese), Acta Mechanica Sinica, Vo1. 1, pp. 48-56.
ELASTIC-PLASTIC CRACK GROWTH ALONG INTERFACES 273

Liu, N. and Drugan, W. J. (1993), Finite Element Solutions of Crack Growth in Incompressible Elastic-
Plastic Solids with Various Yielding Extents and Loadings: Detailed Comparisons with Analytical
Solutions, International Journal of Fracture, Vol. 59, pp. 265-289.
Ponte Castaeda, P. and Mataga, P. A. (1991), Stable Crack Growth Along a Brittle/Ductile Interface. Part 1
Near-Tip Fields, International Journal of Solids and Structures, Vol. 27, 105.
Rice, J. R. (1967), Mechanics of Crack Tip Deformation and Extension by Fatigue, Fatigue Crack
Propagation, ASTM-STP 415, pp. 247-311.
Rice, J. R. (1968), Mathematical Analysis in the Mechanics of Fracture, in Fracture: An Advanced
Treatise, edited by H. Leibowitz, Academic Press, New York, Vol. 2, pp. 191-311.
Rice, J. R. (1974), Elastic-Plastic Models for Stable Crack Growth, in Mechanics and Mechanisms of Crack
Growth, edited by May, M. J., British Steel Corp. Physical Metallurgy Centre Publication, Sheffield, pp.
14-39.
Rice, J. R. (1982), Elastic-Plastic Crack Growth, in Mechanics of Solids: The R. Hill 60th Anniversary
Volume, edited by Hopkins, H. G. and Sewell, M. J., Pergamon Press, Oxford, pp. 539-562.
Rice, J. R. (1987), Tensile Crack Tip Fields in Elastic-Ideally Plastic Crystals, Mechanics of Materials, Vol.
6, pp. 317-335.
Rice, J. R., Drugan, W. J. and Sham, T-L. (1980), Elastic-Plastic Analysis of Growing Cracks, Fracture
Mechanics, ASTM-STP 700, pp. 189-219.
Rice, J. R. and Tracey, D. M. (1973), Computational Fracture Mechanics, in Numerical and Computer
Methods in Structural Mechanics, edited by Fenves, S. J. et al., Academic Press, New York, pp. 585
623.
Sham, T.-L., Li, J. and Hancock, J. W. (1999), A Family of Plane Strain Crack Tip Stress Fields for Interface
Cracks in Strength-Mismatched Elastic-Perfectly Plastic Solids, Journal of the Mechanics and Physics
of Solids, Vol. 47, pp. 1963-2010.
Slepyan, L. I. (1974), Growing Crack During Plane Deformation of an Elastic-Plastic Body, Mekhanika
Tverdogo TeIa, Vol. 9, pp. 57-67.

Appendix: Derivation of Logarithmically Singular Velocity Fields in a General


Propagating Plane Strain Near-Tip Constant Stress Plastic Sector

Here we provide an asymptotic leading-order analysis of the velocity fields in a


propagating constant stress plastic sector with an arbitrary (constant) stress state, i.e.
one of the form (42) satisfying (37). First, employing (42), (6) simplifies to:

(Al)

To leading order as r 0, this reduces to:

(A2)

which for the sake of the upcoming analysis we rewrite as:

(A3a, b, c)
274 W. J. DRUGAN

Employing (2), (A3a, b) are:

(A4a, b)

These may be combined to give a single partial differential equation for v1 :

(A5a, b)

The general solution to (A5a) is:

(A6)

and form this and (A4a) we find the general solution for v2 :

(A7)

In these solutions, and g are arbitrary functions of the indicated variables, and we
have defined

(A8)

Finally, since we seek solutions with velocity fields that are logarithmically singular in
r as r 0, (A6) and (A7) become:

(A9a)

(A9b)

where B 1 and B2 are undetermined constants. From (A3c) and (A9a),

(A10)
STUDY OF CRACK DYNAMICS USING THE VIRTUAL
INTERNAL BOND METHOD

PATRICK A. KLEIN
Sandia National Laboratory
HUAJIAN GAO
Stanford University

It is a great honor for us to contribute to this 60th Anniversary Volume in


honor of Professor James R. Rice and to commemorate his profound
contributions to the science of fracture. All of us are indebted to him for
the leadership he has provided through the excellence of his scholarship as
well as the warm and gentle way he has influenced all his students.
H.G., 1999

Abstract. Most existing theories of fracture are based on small deformation constitutive
models. These approaches are in contrast to the fact that extraordinarily large, nonlin-
ear elastic deformations inevitably occur during brittle fracture. Though the classical
approaches have proved successful in a wide range of applications, they may be inap-
plicable for or have proved incapable of explaining experimental observations in which
nonlinear, hyperelastic material response is an essential feature of the phenomenon. The
fracture path instabilities observed during dynamic propagation of cracks are among
the phenomena that have not yet been thoroughly explained. Simulation approaches
that incorporate a cohesive view of material are able to demonstrate the appearance of
crack tip instabilities. Molecular dynamics and cohesive surface methods are among the
methods that exhibit this behavior. A virtual internal bond (VIB) model with randomized
cohesive interactions between material particles has been proposed as an integration of
continuum models with cohesive surfaces and atomistic models with interatomic bonding.
This approach differs from an atomistic model in that a phenomenological cohesive force
law is assumed to act between material particles which are not necessarily atoms. It
also differs from a cohesive surface model in that, rather than imposing a cohesive law
along a prescribed set of discrete surfaces, a randomized network of cohesive bonds is
statistically incorporated into the constitutive response of the material via the Cauchy-
Born rule, by equating the strain energy function on the continuum level to the potential
energy stored in the cohesive bonds due to an imposed deformation. The approach could
be viewed as an attempt to provide a more physical basis for the hyperelastic constitutive
laws used in finite strain continuum mechanics. This approach allows the phenomenon
of dynamic crack tip instabilities to be analyzed within a continuum framework. Direct
simulation of crack growth without a presumed nucleation, growth, or branching criterion
lends support to the theory that instabilities occur as a result of a local limiting speed,
governed by the rate at which bond-breaking information can be transmitted to the
material ahead of a propagating crack.
275
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 275309.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
276 P. A. KLEIN AND H. GAO

1. Introduction

Cracks propagating at high speeds in brittle materials display behavior that


the fracture community has yet to find consensus in explaining. Much of
the current debate involves the factors that determine the terminal velocity
of propagating cracks and the roughened appearance of fracture surfaces
produced under fast fracture. The experimental methods used to study
dynamic fracture, some key experimental observations, as well as a review
of a number of models that attempt to explain the observed phenomena are
reviewed by Ravi-Chandar (Ravi-Chandar, 1998). The theoretical limiting
speed for crack propagation is the Rayleigh surface wave speed c R . How-
ever, experimental studies of cracks propagating through brittle materials
typically quote terminal crack speeds of approximately c R /2, while insta-
bilities have been observed for crack speeds as low as cR /3. Detailed experi-
ments of dynamic crack propagation through the brittle polymer Homalite-
100 appear in the studies of Ravi-Chandar and Knauss (Ravi-Chandar
and Knauss, 1984a; Ravi-Chandar and Knauss, 1984b; Ravi-Chandar and
Knauss, 1984c; Ravi-Chandar and Knauss, 1984d). They observe the well-
known mirror-mist-hackle appearance of the fracture surface. An image
from their study (Ravi-Chandar and Knauss, 1984b) showing the fracture
surface is reproduced in Figure 1.

Figure 1. Mirror-mist-hackle appearance of a fracture surface observed by


Ravi-Chandar and Knauss (Ravi-Chandar and Knauss, 1984a; Ravi-Chandar and
Knauss, 1984b; Ravi-Chandar and Knauss, 1984c; Ravi-Chandar and Knauss, 1984d).

The direction of crack propagation is from left to right in the figure.


The image shows that the crack surface becomes progressively rougher as
the crack extends and the crack speed gets correspondingly higher. By the
method of caustics, they are able to observe distinctly different behavior
CRACK DYNAMICS USING VIB METHOD 277

in the propagating crack front that helps explain the appearance of the
fracture surface. In the initial stages of growth, the crack extends by the
propagation of a single, curved front, similar to what is observed under
quasi-static conditions. The fracture surface appears essentially smooth,
leading to its description as the mirror zone. As growth progresses, the
crack front appears to extend by the action of multiple, smaller crack fronts.
The opaque appearance of the crack surface due to fine-scale roughness
leads to its description as the mist zone. In the hackle zone, the scale
of the multiple crack fronts increases, leading to a larger degree of fracture
surface roughening.
The emergence of crack tip instabilities under dynamic conditions was
first predicted analytically in a study by Yoffe (Yoffe, 1951). Her analysis
predicts branching at 60 when the propagation speed reaches approxi-
mately 0.6 c R . Though the result is significant in demonstrating the onset
of fracture path instabilities, neither the crack speed nor the branching an-
gle agrees well with experimental observations. Branching angles typically
fall between 10 and 45, while instabilities appear at speeds well below
Yoffes predicted value. In particular, Fineberg and coworkers (Fineberg
et al., 1991; Fineberg et al., 1992) make detailed measurements of crack
propagation in PMMA, showing the onset of oscillatory behavior in the
crack tip at speeds of roughly c R /3. Classical theories are unable to ex-
plain the development of crack tip instabilities at these speeds. It has been
suggested by Gao (Gao, 1996; Gao, 1997) that the hyperelastic nature of
crack tip deformation may be essential for understanding dynamic crack tip
instabilities. Elastic softening due to hyperelastic stretching can drastically
change the way acoustic waves propagate in a solid. The variation in the
elastic properties with deformation referred to in this setting should not
be associated with models of continuum damage. Damage models attempt
to incorporate the effects of microstructural evolution in the behavior of a
material, such as microcrack distributions (Kachanov, 1992) or the nucle-
ation and coalescence of voids (Gurson, 1977). The elastic softening under
consideration in the present context is an effect of the continuously varying
stiffness of a nonlinear cohesive interaction. To illustrate this point, let us
consider how a long wavelength signal propagates along an atomic chain
stretched toward its cohesive limit. Before the chain is stretched, it behaves
like an elastic bar along which a longitudinal wave can be transmitted
with speed equal to where E denotes the initial stiffness (Youngs
modulus) and the density of the chain. As the chain is stretched near its
cohesive (fracture) limit, the tangent stiffness vanishes while a finite tension
builds up with magnitude equal to the cohesive stress Tmax . At this point,
the chain behaves like an elastic string along which a transverse wave can
be transmitted with speed equal to Note that the cohesive-state
278 P. A. KLEIN AND H. GAO

wave speed depends on the magnitude of the stress rather than the slope
of the stress-strain curve. During this hyperelastic deformation, the slope
of the force-displacement (or stress-strain) curve continuously decreases
from its initial value corresponding to the elastic stiffness of the chain, and
ultimately vanishes at the cohesive limit. Clearly, a linear theory assuming
constant stiffness cannot be used to investigate the behavior of the chain
near its failure point.
Generalizing the above concept, Gao (Gao, 1996; Gao, 1997) has shown
that a homogeneous, isotropic, hyperelastic solid exhibits this string-like be-
havior near its plane strain equibiaxial cohesive stress in that the cohesive-
state wave speed is equal to where max is the cohesive stress and
is the density of the undeformed solid. The state of plane strain equibiaxial
stress resembles the condition that a material particle experiences in front
of a mode I crack tip. This observation comes directly from Irwins (Irwin,
1957) classic crack tip fields. This cohesive state is identified as the bottle-
neck state (Gao, 1996) for transmission of fract ure signals ahead of a mode
I crack tip. The crack propagation velocity is limited by how fast elastic
waves can transport strain energy ahead of the crack to sustain the bond
breaking processes in the fracture process zone. From this point of view,
the cohesive-state wave speed leads to the concept of local limiting fracture
speed (Gao, 1996) which has provided an explanation for the mirror-mist-
hackle instabilities. Assuming values of max= E /30 and a Poissons ratio
of v = 1/4, the cohesive state wave speed is roughly 0.32 c R , which is in good
agreement with Finebergs measurements. Gaos (Gao, 1993) wavy crack
model presents a multiscale view of a dynamic crack in which the global, or
apparent, crack speed is driven near c R /2 to maximize the fracture energy
absorbed by the advancing crack, while the microscopic tip speed may be
significantly higher in response to local crack driving forces. A multiscale
view may also be helpful in understanding the mirror-mist-hackle insta-
bilities as the crack path begins to deviate in an attempt to circumvent
the acoustic barrier presented by elastically softened material immediately
ahead of the crack. Hyperelastic constitutive models that are capable of
describing a materials transition to fracture have not been available for
studying these effects. As will be shown, the Virtual Internal Bond (VIB)
model displays this cohesive state behavior and should therefore be capable
of reproducing dynamic crack tip instabilities.
The atomistic simulations of Abraham and coworkers (Abraham et al.,
1994; Abraham, 1997) display dynamic crack instabilities reminiscent of the
experimental results though at decidedly smaller length and time scales.
The numerical studies of Xu and Needleman (Xu and Needleman, 1994),
employing networks of cohesive surface elements, are also successful in pro-
ducing crack tip instabilities that qualitatively match the experimentally
CRACK DYNAMICS USING VIB METHOD 279

observed behavior. In their study, the continuum elements are modeled


using the Kirchhoff-St.Venant constitutive relations, that do not display any
softening, while the cohesive surfaces are governed by a traction-separation
potential with qualitative features similar to those discussed for the poten-
tial in Section 5. Their observation of instabilities may also be explained
with the local limiting speed theory since the composite material com-
posed of continuum elements with intervening cohesive surfaces has the
effective properties of a cohesive continuum, though with mesh dependent
anisotropic characteristics.

2. The Virtual Internal Bond model

The general features of the VIB model (Gao and Klein, 1998), as well as
more detailed study of its localization behavior (Klein and Gao, 1998; Klein,
1999), are described elsewhere. A brief review of the model is repeated
here for completeness. The model attempts to incorporate the behavior of
a spatial distribution of cohesive bonds that are presumed to act at the
microstructural level. The link between the microstructure and continuum-
level measures of deformation is made using the Cauchy-Born rule. This
approach is employed by Ortiz, Phillips, and Tadmor to develop constitutive
models for single crystal materials (Tadmor et al., 1996). From a practi-
cal point of view, it is desirable to extend the procedure for embedding
cohesive behavior beyond models for single crystal materials. We would
like to develop a model to study fracture in noncrystalline materials, such
as polymers or glasses, or in polycrystalline materials for which the elastic
anisotropy is lost due to the homogenizing effect of many randomly oriented
crystalline grains. The concept of homogenization is applied here to the
interactions between particles. The resulting constitutive model embeds the
cohesive nature of the interaction potentials, but is free of any orientational
dependencies.
The VIB model is developed within the framework of hyperelasticity.
The current, or deformed, configuration of a body is described as x = (X),
by a mapping of the undeformed configuration X. The arrangement of
cohesive interactions among material particles is described by a spatial bond
density function. The strain energy density is computed by integrating the
bond density in space in a continuous analog to the sum over discrete lattice
neighbors for the case of crystalline materials. The VIB form of the strain
energy density function is

(1)
280 P. A. KLEIN AND H. GAO

where is the undeformed representative volume, l is the deformed virtual


bond length, U(l) is the bonding potential, is the volumetric bond
density function, and is the integration volume defined by the range
of influence of U. Depending on the range of influence of the bond po-
tential function, the integration volume may not correspond with the
representative volume . This difference may be illustrated for crystalline
materials whenever the bond potentials extend beyond the lattice unit cell.
This method was first alluded to by Gao (Gao, 1996) as a method for con-
structing an amorphous network of cohesive bonds by a spatial average. The
deformed bond length l is computed from the Cauchy-Born rule, assuming
the integration volume deforms homogeneously as described by a given
deformation gradient In order to avoid questions as to whether the
Cauchy-Born rule holds for the proposed microstructure, we consider only
bond density functions that are centrosymmetric. Under this restriction,
the deformation at the microstructural level must be homogeneous in order
to maintain the symmetry present in the undeformed configuration. In
extending this description from a lattice of discrete bonding interactions
to a continuous bond distribution, the invariance in the deformation with
respect to translation of our observation point is not limited to positioning
the coordinate origin at crystal lattice sites. All points in the material act as
centers of symmetry. From physical considerations, we assume the bonding
function U will be relatively short range so that, although the bond density
function describes the relative particle distribution of the entire body
about the observation point, the integration volume will be on the order
of the representative volume The undeformed virtual bond vector is
represented as

(2)

where L is the reference bond length, and is a unit vector in the direction
of the undeformed bond. Undeformed bonds are mapped to their deformed
configuration l by

(3)

Making use of the right Cauchy-Green stretch tensor, the deformed bond
length is

(4)

Expressed as a function of C, the deformed bond length transforms objec-


tively by construction. The stress response and tangent moduli are com-
puted from the strain energy density (1) using the relations from Green
CRACK DYNAMICS USING VIB METHOD 281

elastic theory (Marsden and Hughes, 1983). Employing

(5)

the (symmetric) 2nd Piola-Kirchhoff stress is

(6)

and the material tangent modulus is

(7)

The modulus (7) displays Cauchy symmetry as well as the usual major and
minor symmetries of elasticity. This result that an amorphous solid with a
random network of cohesive bonds satisfies the Cauchy relation seems to be
a generalization of Stakgolds theorem (Stakgold, 1950) to an amorphous
solid. A material particle in an amorphous solid satisfies centrosymmetry
in a statistical sense, and the cohesive force law corresponds to a two-body
potential.
Since we limit our study to centrosymmetric bond density functions, it
is natural to express the strain energy density (1) in spherical coordinates
as

(8)

where each bond is characterized by coordinates {L, , }. For conciseness,


we have introduced the notation of a spherical average as

(9)

where L * represents the maximum distance over which particles interact.


In spherical coordinates, the bond direction vector can be written as

(10)

The precise definition of D ( L, , ) is that rep-


resents the number of bonds in the undeformed solid with length between L
282 P. A. KLEIN AND H. GAO

and L + dL and orientation between { , } and { + d , + d }. Using this


notation, the components of the 2nd Piola-Kirchhoff stress and the material
tangent modulus from (6) and (7) can be represented as

(11)

and

(12)

There are a few special cases with regard to the bond density function
D (L, ,):
(1) The case

(13)

corresponds to a network of identical bonds of undeformed length L 0.


The Dirac delta function is denoted here with D . A crystal lattice such
as face-centered cubic with interactions limited to only first nearest
neighbors can be represented as

(14)

where D 0 is a scaling constant.


(2) The case

(15)

represents a transversely isotropic solid. It will be convenient in the


subsequent discussion to define a notation for the spatial average under
the transversely isotropic case as

(16)

(3) The case

(17)
CRACK DYNAMICS USING VIB METHOD 283

yields an isotropic solid with fully randomized internal bonds. We will


shortly see that the instantaneous response of this solid becomes trans-
versely isotropic under equibiaxial stretching. The instantaneous re-
sponse becomes generally anisotropic under finite deformations. The
associated average for the isotropic case is

(18)

Selecting the radial bond distribution function as


(19)
where D 0 is a constant, implies that all of the bonds in the solid are
of the same type, or initial length. The result is a model for amor-
phous material with nearest neighbor bonding only. With this bond
distribution, the strain energy density from (8) and (9) becomes

(20)

(4) The case


(21)
yields a plane stress, isotropic solid. Unlike the standard plane stress
approximation, this model is truly two-dimensional, displaying no out-
of-plane deformation or stress. The material can be thought of as being
composed of a single sheet of cohesive bonds, or layers of noninteracting
sheets, since all of the bonds lie in a single plane. Selecting the radial
bond distribution function as
(22)
where D 0 is a constant, implies that all of the bonds in the solid are of
the same type, or initial length. As was the case in three dimensions,
the result is a model for amorphous material with nearest neighbor
bonding only. For this case, the strain energy density has the especially
simple form

(23)

Though this model may be physically unrealistic, its simplicity makes it


useful for both theoretical and numerical investigations of the fracture
properties resulting from the VIB approach.
284 P. A. KLEIN AND H. GAO

In the sections that follow, we will investigate the stress response and
fracture properties exhibited by the VIB model. A majority of the discus-
sion will center on the isotropic model defined above for three and two
dimensions by cases (3) and (4), respectively.

3. Isotropically elastic properties at infinitesimal strain

We are particularly interested in the properties of an amorphous network


of cohesive bonds that is homogeneous and isotropic at small strain. The
properties of the model at small strains can be used to fit the model pa-
rameters to particular materials. For this case, the bond density function is
selected as D ( L,,, ) = D L (L ), and E, the Green Lagrangian strain, and
S reduce to the strain and stress tensors of linear elasticity: E a n d
S . can be expanded up to the quadratic term as

(24)

where the first order term does not appear since the interaction potential
is assumed to satisfy U '( L) = 0 to give a stress-free, undeformed state.
This produces the generalized Hookes law
(25)
where the modulus under infinitesimal deformations is related to the cohe-
sive bond distribution by

(26)

It can be directly verified that

(27)

is a fourth-rank isotropic tensor. The resulting elastic stiffness tensor is

(28)
with shear modulus equal to

(29)
CRACK DYNAMICS USING VIB METHOD 285

The other elastic constants are

(30)

where and are the Lam constants, v is Poissons ratio, and E is Youngs
modulus. Since a variety of engineering materials display values of Poissons
ratio within a range around l/4, the Cauchy symmetry exhibited by the
VIB model does not represent a significant restriction of the formulation
for application to practical analyses.
The results for the plane stress, isotropic bond density function
(21)
are quite similar, but the relations for the elastic moduli (29,30) must be
modified to account for the difference in spatial dimensions. The modulus
in the generalized Hookes law (25) is given by

(31)

The elastic stiffness tensor has the same form given in (28). However, the
shear modulus becomes

(32)

From Cauchy symmetry, the Lam constants are equal ( = ), but the
relations between the other elastic constants are different from the standard
relations of three dimensional elasticity. From the Hookes law relations in
strictly two dimensions, Poissons ratio, Youngs modulus, and the bulk
modulus are related to the Lam constants by

and (33)

Using (33), the elastic constants for the plane stress isotropic VIB model
are

and (34)

4. Elastic properties under plane strain equibiaxial stretching

The state of plane strain, equibiaxial stretching resembles the deformation


a material particle experiences ahead of a mode I crack tip. The elastic
286 P. A. KLEIN AND H. GAO

properties under such a state of deformation at the cohesive limit of the


material have been of key importance in the theory of local limiting fracture
speeds (Gao, 1996; Gao, 1997).
For plane strain equibiaxial stretching, the deformation gradient is

(35)

Under this deformation, the bond length from (4),


(36)

becomes independent of the in-plane orientation angle . The components


of the 2 nd Piola-Kirchhoff stress tensor can be calculated from (11) as

(37)

The components of the Cauchy stress are

(38)
CRACK DYNAMICS USING VIB METHOD 287

Note that the equibiaxial stretching results in equibiaxial stress. This is not
necessarily true for a crystalline solid and indicates that our representation
of an amorphous network of cohesive bonds does respond isotropically.
The nonzero components of the material tangent modulus are obtained
from (12) as

(39)
and from the Cauchy relations

where

(40)

The components of the elastic modulus in (39) exhibit the symmetry


of a transversely isotropic material. Therefore, under equibiaxial stretching
the initially isotropic solid develops a transversely isotropic instantaneous
response.
When the equibiaxial stress reaches the cohesive limit max , the condi-
tion

(41)
288 P. A. KLEIN AND H. GAO

or, equivalently

(42)

is satisfied. The following relations can be observed from (39):

(43)

which when combined with (42) indicate that all the in-plane components
of C I J K L vanish,

for (44)

In order to understand the implications of these results, we introduce


some additional concepts from the theory of wave propagation. In La-
grangian coordinates, the equation of elastodynamics at finite deformations
is

(45)

where P = FS is the (nonsymmetric) 1st Piola-Kirchhoff stress tensor and


0 is the density of the undeformed material. A plane wave propagating
through the continuum is described by

(46)

where u is the displacement of a material particle, k is the wave number, i


is , c is the wave speed, and N is the propagation direction defined in
the reference configuration. Inserting the plane wave description (46) into
the equation of motion (45) yields

(47)

where

(48)

is the acoustical tensor. Nontrivial solutions of (47) require that the deter-
minant of the coefficient matrix of a vanish, yielding the characteristic bulk
wave speeds from the eigenvalues of q(N). The effective modulus B in (48)
can be expressed in terms of the Lagrangian representation of stress and
modulus as

(49)
CRACK DYNAMICS USING VIB METHOD 289

Using the results for the stress and modulus at the cohesive limit (38,39,44),
the speed of elastic waves propagating in the (X 1 , X 2 )-plane is

(50)

This result agrees with the previous analysis of Gao (Gao, 1996; Gao, 1997)
but reaches the same conclusion from a different perspective. His analysis
only assumes that the material possesses general cohesive properties and is
not derived from the description of a particular constitutive model. Note
that waves can travel with a different speed in an out-of-plane direction.
This issue is not addressed here. The ratio

(51)

is the plane strain Poissons ratio at finite deformation. Gao (Gao, 1997)
has shown that the cohesive state wave speed is given by (50) as long as
Poissons ratio v e does not approach 1 as the cohesive stress is reached. In
the present case, v e equals l/3 irrespective of the magnitude of the strain.
The cohesive state wave speed is intimately connected to the concept of
local limiting fracture speed which has been proposed as an explanation for
the occurrence of the mirror-mist-hackle dynamic crack tip instabilities.
The discrepancy between experimental observations of this speed and the
established theory of dynamic fracture (Freund, 1990) has been discussed in
previous work by Gao (Gao, 1993; Gao, 1996). A close examination of the
mode I crack tip deformation suggests (Gao, 1996) that the state of plane
strain, equibiaxial cohesive stress is the bottleneck state for transmission of
bond breaking fracture signals ahead of a mode I crack. The crack propaga-
tion velocity is limited by how fast elastic waves can transport strain energy
to material ahead of the crack tip to sustain the bond breaking processes
in the fracture process zone. Prom this point of view, the cohesive state
wave speed leads to the concept of local limiting fracture speed at which a
straightforward propagation of the crack becomes unfavorable. The crack
then attempts to choose an inclined growth direction through material
with less elastic softening. Along an inclined direction, stress waves can
propagate faster than in the straight-ahead direction where elastic softening
has caused all of the in-plane moduli to vanish. As discussed by Gao (Gao,
1993; Gao, 1996), from a global point of view, the apparent crack motion is
in a relatively low inertia state when the local crack branching occurs. As
a result of the local-global inertia competition, the crack chooses to propa-
gate along a wavy path, thus exhibiting the mirror-mist-hackle instability.
Unlike other descriptions of fracture, this explanation incorporates not only
the role of tractions generated as a results of crack opening displacements,
290 P. A. KLEIN AND H. GAO

but also the effect of deformation-induced variations in material response


along the propagation direction. Taking some typical values for the cohesive
strength relative to Youngs modulus and Poissons ratio as max = E /30
and v = l/4, we estimate that

(52)

where c R is the Rayleigh surface wave speed. This is precisely the speed
at which crack tip instability is observed experimentally (Fineberg et al.,
1991).
The present calculation with amorphous cohesive bonds provides further
support for the theory of local limiting fracture speed. It also shows that by
introducing the essential features of a cohesive view of material, the present
model exhibits the correct cohesive state wave speed and should be suitable
for use in numerical simulations of dynamic fracture.

5. A model interaction potential

The preceding discussions have not made reference to any particular co-
hesive potential. For the purpose of demonstration, we introduce the phe-
nomenological cohesive force law

(53)

The defining characteristics of this potential are illustrated in Figure 2.


The potential has a well of depth U 0 = U (L) = A B 2 . This quantity
can be related to the fracture energy through a J-integral analysis (Klein
and Gao, 1998). The response of the potential for bond lengths around the
location of this minimum is determined by the parameter A. The stiffness
at the minimum is U "(L) = A . In the left portion of Figure 2, the potential

Figure 2. Characteristics of the phenomenological cohesive force potential.


CRACK DYNAMICS USING VIB METHOD 291

U is shown with a quadratic potential exhibiting the same energy, force,


and stiffness at l = L. The quadratic potential defines the response of a
linear spring, for which the force always varies proportionally with the
stretch in the bond. Compared with the quadratic potential, the cohesive
potential displays greater stiffness for compressive deformations and the
characteristic loss of convexity in tension.
The second parameter B determines the cohesive properties of the po-
tential. The maximum force is generated for a stretched length
of In terms of the 2nd Piola-Kirchhoff stress (6), we see that
the cohesive strength displayed by the VIB model is not determined by the
force U (l ), but rather by the ratio U (l ) / l . The maximum value for this
ratio occurs for

(54)

from which we find

(55)

6. Failure indicators

The promise of cohesive modeling is the ability to produce fracture in nu-


merical simulations without requiring an imposed failure criterion. Failure
occurs as a natural consequence of the cohesive formulation, which embeds
a finite strength and fracture energy in the material model. Although we
are not required to impose a failure criterion, we do need to identify a
failure indicator. With classical fracture theory, conditions for crack prop-
agation are posed as a threshold criterion that clearly defines the point
at which failure occurs. With a cohesive view of material, failure proceeds
in a much more gradual manner. The material does not simply fail, but
exhibits a complete history of response from the undeformed state through
the cohesive limit to complete failure.
In selecting a failure indicator, we choose to associate it with a threshold
condition that corresponds to reaching the cohesive limit rather than a cut-
off condition when the stresses being sustained by the material are deemed
insignificant. Hill (Hill, 1962) has described the loss of strong ellipticity of
the strain energy density function as an indication of the loss of stability of
a solid. The analysis is related to the eigenvalues of the acoustical tensor
q(N) (48). The loss of strong ellipticity coincides with a loss of uniqueness
in the solutions of the governing equation of elastodynamics (45). The
discontinuous modes of deformation that become admissible with the loss of
292 P. A. KLEIN AND H. GAO

strong ellipticity appear across a characteristic surface, which we describe


with the normal N * in the undeformed configuration. With the loss of
strong ellipticity produced by models of plasticity exhibiting softening,
these characteristic surfaces are associated with slip, or shear, bands. In
the present context, we associate these characteristic surfaces with highly
localized bands of deformation that become new crack faces as the material
reaches complete failure.
We also study the behavior of a second failure indicator that is related
more directly with the cohesive strength than is the analysis of the acous-
tical tensor. Since the bond density functions of the VIB model are defined
with respect to the undeformed configuration, the 1s t Piola-Kirchhoff stress
P is related to the force generated by the network of virtual bonds. The
maximum values of P corresponds to the largest forces that can be sustained
by the bonds of the virtual microstructure.
The onset of failure indicated by these two criteria as a function of the
state of deformation is shown in Figure 3 for the plane stress, isotropic
VIB model. The configuration of the imposed deformation is shown in

Figure 3. (a) Geometry of the loading configuration and (b) a comparison of fail-
ure indicators for the plane stress, isotropic VIB model as a function of the state of
deformation.

Figure 3(a). The deformation is characterized by E 1 / E2 , the ratio of the


principal values of the Green strain, where E 1 and E2 are the strains
along and perpendicular to the primary loading direction, respectively.
The normal to the characteristic surface associated with the loss of strong
ellipticity is shown as N*. Figure 3(b) illustrates how the two criteria differ
in indicating the onset of failure as a function of the state of deformation
for B / L = 0.05 using the model interaction potential from Section 5. The
CRACK DYNAMICS USING VIB METHOD 293

stretch in the primary loading direction at the point of failure is designated


as *. The deformation states for uniaxial stress (P2 = 0) and uniaxial strain
(E2 = 0) are indicated in the figure. For E 2 / E 1 < 0.3, the loss of strong
ellipticity clearly occurs at a smaller stretch than the stretch required to
reach the maximum value of P 1 . In this regime, the maximum value of
P 1 attained at * in the figure is unreachable since homogeneous states
of deformation become unstable once the acoustical tensor is non-positive
definite. As the imposed deformation becomes more nearly equibiaxial, the
critical stretches for the two criteria converge. For E2 / E 1 > 0.3, the two
curves in the figure are nearly indistinguishable. For the equibiaxial case,
both criteria indicate the onset of failure at * = B /L, the stretch for which
the bond force is maximal.
Figure 4 shows the orientation of the characteristic surface with normal
N* as a function of the state of deformation. The angle * is measured
from the axis of primary loading, as indicated in Figure 3(a). States of

Figure 4. Orientation of the localized band as a function of the deformation.

deformation corresponding to uniaxial stress and strain are indicated in the


figure. The key result of these calculations is that the characteristic surface
associated with the loss of strong ellipticity is oriented perpendicular to
the primary loading direction for imposed deformations approaching the
equibiaxial state, E 2 / E1 > 0.5 in the figure. From the asymptotic crack tip
solutions of linear elasticity, we know that the state of equibiaxial stretching
resembles the deformation a material particle experiences in front of a mode
I crack tip. Therefore, it is encouraging that the direction of crack growth
indicated by N* matches our expectations for the direction of propagation
294 P. A. KLEIN AND H. GAO

of a mode I crack. Under mixed-mode conditions, classical fracture theory


requires that we postulate a crack propagation direction. Typically, one
prescribes crack propagation in the locally mode I direction, the orien-
tation for which K I I = 0. Alternatively, one could select the direction
corresponding to the largest hoop stress surrounding the crack tip.
Since the mechanisms of material failure are not well understood, the most
appropriate choice for the direction of crack propagation is also unclear.
With the VIB model, the fracture characteristics under mixed-mode
conditions is embedded in the constitutive behavior. Once the parameters
in model have been selected, no additional criteria need to be imposed
in order to reproduce a wide variety of fracture phenomena. As shown in
Figures 3 and 4, the loss of strong ellipticity appears to be an appropriate
indicator of the onset of failure. For loading with mode I character, it agrees
with our expectation for failure to initiate as the stress in the material
approaches the cohesive limit and for crack propagation to occur in the
direction perpendicular to the direction of the largest stress.

7. Numerical simulations

In order to demonstrate the capability of the VIB model to reproduce dy-


namic crack tip instabilities, we present two-dimensional simulations using
the plane stress, isotropic VIB model and three-dimensional simulations
using the VIB model in principal stretches (Klein, 1999). Details of the VIB
model in principal stretches will not be described here, but we note that
expressing the model in principal stretches greatly improves computational
efficiency for three-dimensional problems. For these calculations, the VIB
model is used within an updated Lagrangian finite element formulation. The
equations of motion are integrated in time using an explicit, central differ-
ence scheme from the classical Newmark family of methods. The geometry
of the finite element models is shown in Figure 5. As noted in the figure, the
three-dimensional model possesses considerably more degrees of freedom
although its overall dimensions are smaller. The two-dimensional model
contains 159,175 nodes in 158,705 elements for a total of 318,106 degrees
of freedom, accounting for the nodes with prescribed kinematic bound-
ary conditions. The three-dimensional model contains 1,305,702 nodes in
1,260,000 elements for a total of 3,865,800 degrees of freedom. The models
are loaded by symmetrically prescribed velocity boundary conditions for
the nodes along the upper and lower edges of the domains. The bound-
ary nodes are accelerated to the prescribed velocity over a time of 0.2 s,
essentially simulating an impact load. Additional boundary conditions are
prescribed to prevent rigid body motion. For the three-dimensional simu-
lations, the remaining surfaces are traction-free. The material parameters
CRACK DYNAMICS USING VIB METHOD 295

Figure 5. Geometry and dimensions of the two- and three-dimensional finite element
models used to study dynamic crack tip instabilities. The size of the three-dimensional
model is shown on the two-dimensional model with a dashed outline.

are selected to be representative of PMMA, the material used in the ex-


periments of Fineberg (Fineberg et al., 1991; Fineberg et al., 1992), with
E = 3.24GPa, v constrained by Cauchy symmetry as described in Sec-
tion 3, and = 1200 kg/m3 . These values produce dilatational and shear
wave speeds of cd = 1740 m/s and c s = 1000 m/s, respectively, and a
296 P. A. KLEIN AND H. GAO

Rayleigh wave speed (Freund, 1990) given by

(56)

The fracture energy is selected as 350 J/m2 , and the cohesive strength is
selected as E /30. A J-integral analysis can be used to show that the fracture
energy depends on mesh size with the current form of the VIB model (Klein
and Gao, 1998). The given cohesive parameters dictate that elements in
the central region of height 1.0 mm have dimension of h = 9.2 m. A three-
parameter potential could be created to allow the elastic properties, fracture
energy, and cohesive strength to be selected independently. However, this
approach represents only a partial solution to the well-known difficulties
associated with simulating strain localization. For the two-dimensional sim-
ulations, four-noded quadrilateral elements are used throughout the model.
The central region is meshed with a regular arrangement of square elements
with dimension 9.2 m while the elements in the outer region increase in
size with distance away from the central zone. For the three-dimensional
simulations, the entire domain is discretized into a regular, structured grid
of eight-noded hexahedral bricks with dimension 9.2 m.
A pre-crack of length 0.5mm ensures that propagation is initially di-
rected along the centerline of the model, though its path is not prescribed
in any way. The simulation results show that crack growth, instabilities,
and branching emerge naturally from the properties of the VIB constitu-
tive model. For the two-dimensional simulations, the apparent length of
the crack is monitored by tracking the elements in which the acoustical
tensor (48) is no longer positive definite. Checking this condition requires
searching all wave propagation directions to see if there are any for which
the wave speeds vanish. Since the models are constrained to prevent rigid
body translations, the crack length is taken as the greatest distance along
X1 between the original crack tip position and the centroids of localized
elements in the undeformed configuration. The time step for the explicit in-
tegration scheme is selected so that t cd / h = 1/3, where h is the minimum
element size.
With the current state of our simulation procedures, the two-dimension-
al simulations are better-suited to quantitative study than are the three-
dimensional simulations. Working in two dimensions, we are able to make
the computational domains larger, reducing the effect of boundaries on the
crack behavior, and analysis of the acoustical tensor requires considerably
less effort. For these reasons, quantitative analysis of the two-dimensional
simulations is presented, while the results of the three-dimensional sim-
ulations are more qualitative. We present results of the two-dimensional
simulations for three values of the prescribed boundary velocity, BC =
CRACK DYNAMICS USING VIB METHOD 297

{2.6,5.2,10.4} m/s. The crack length as a function of the simulation time


for all three cases is shown in Figure 6. The time to the initiation of crack

Figure 6. Apparent crack length over time for three different values of the imposed
boundary velocity.

growth decreases as the impact velocity increases, presumably in response


to the time required to generate the critical driving force at the initial crack
tip. For the impact velocities of 2.6m/s and 5.2m/s, the figure shows that
the crack moves forward and then stops before accelerating to the terminal
velocity. Based on the dimensions of the model and the dilatational wave
speed in the material, the initial loading wave reaches the crack tip from the
boundaries after roughly 1.35 s. An average terminal velocity is calculated
for each case from the slope of the crack length curves a ( t ) beyond the
initial transient behavior. The terminal velocity increases from 0.48cR to
0.53 cR as the impact velocity increases from 2.6 to 10.4m/s.
Figure 7 shows the crack morphologies for the three cases. Each point
along the fracture path marks an element for which the acoustical tensor
condition for localization is satisfied at one or more of the element integra-
tion points. The images in the figure actually correspond to superpositions
of the fracture path history over the entire simulation time. Since the VIB
model is entirely elastic and does not incorporate any irreversibility in the
fracture processes, secondary branches along the fracture path heal after
the leading edge of the crack has advanced far enough to unload the material
in its wake. The fracture paths indicate that the cracks initially propagate
straight ahead along the symmetry line of the domain. In each case, the
first deviation from straightforward propagation is marked by a symmetric
298 P. A. KLEIN AND H. GAO

Figure 7. Fracture patterns for impact velocities of (a) BC = 2.6m/s, (b)


BC =5.2m/s, and (c) BC = 10.4m/s.

branch at roughly 10-15 to the initial crack plane. This first branch occurs
sooner in time and at a shorter crack length as the impact velocity is
increased. The subsequent fracture path becomes more irregular as one
of the two branches arrests while the other continues to propagate. Once
the symmetry of the original propagation has been disrupted, the branches
become more irregular. In both Figures 7(b) and (c), the branching angles
become noticeably larger as the crack advances, reaching a maximum angle
of approximately 35 in (b), and reaching almost 55 in (c).
The initial, straight ahead propagation of the crack is mirror-like,
while the branching at later times forms a hackle zone. What is not
evident from the fracture paths shown in Figure 7 is that some indication
of instability appears ahead of the crack tip significantly earlier than the
occurrence of the first branch, representing the mist mode of propagation.
Figure 8 shows the arrangement of elements in which the acoustical tensor
condition is met at four instants leading to the first branch in the fracture
path for the case of BC = 5.2 m/s. Initially (a), all elements displaying
localization lie along a straight path extending from the pre-crack. After
5.1 s (b), the first evidence of localization in elements above and below
the symmetry line appears. Based on the local limiting speed theory of
dynamic crack tip instabilities, the crack has reached a speed at which
the strain softened material immediately ahead of the crack is unable to
maintain a sufficient rate of energy transfer, and the crack has begun to
probe alternate propagation directions. Between 5.1 and 5.9 s after impact
(c), the crack continues to accelerate and the acoustical barrier ahead of the
CRACK DYNAMICS USING VIB METHOD 299

Figure 8. The onset of branching for B C = 5.2m/s.

crack tip enlarges, evolving to an extended region of damaged material.


Since deformations in the VIB model are strictly reversible, the material
recovers as the tip moves away, leaving no indication of this extended region
in the subsequent fracture path. At some time before 6.1 s (d), the crack tip
reaches a critical state, and the first true branch appears in the crack path.
The sequence of Figures 8(a)-(d) bears resemblance to the mirror-mist-
hackle progression of crack face roughness observed in experiments. The
small scale roughness in the numerical results is too fine to be resolved with
the current mesh dimensions, but the transition to larger scale roughness
is clearly displayed.
The crack length data in Figure 6 can be numerically differentiated in
order to calculate the apparent crack velocity as a function of time. The
crack velocity, normalized by the Rayleigh wave speed cR , for BC = 5.2 m/s
is shown in Figure 9. As is evident from the crack length data, the crack
moves forward at approximately 2.0 s after the impact occurs, arrests, and
then accelerates quickly to an average terminal velocity of approximately
0.5 c R . As observed by Fineberg (Fineberg et al., 1992), the crack acceler-
ates quickly, but continuously, instead of initiating at the terminal velocity.
As the crack accelerates and rapid branching begins, the crack velocity
becomes more irregular. This result qualitatively matches the experimental
measurements as well as the numerical results obtained by Xu and Needle-
man (Xu and Needleman, 1994). Markers A-D indicate the times at which
the crack is pictured in Figures 10-13. The four points correspond to (A)
initial stages of propagation in the straight-ahead direction; (B) the point
at which the small scale instabilities shown in Figure 8(b) appear; (C) the
point shown in Figure 8(d) at which the first clear branch appears; and (D)
propagation at the terminal velocity.
Figures 10-13 show two views of the propagating crack at the four
points indicated in Figure 9. The upper plot (a) in each figure shows the
300 P. A. KLEIN AND H. GAO

Figure 9. Apparent crack velocity over time for an impact speed of BC = 5.2 m/s
with markers A-D indicating the times at which the crack is pictured in Figures 10-13.

distribution of the instantaneous shear wave speed cs for a wave traveling


in the X 1 -direction. The contours are normalized by the shear wave speed
in the undeformed material (c s ) 0 . These figures are intended to show how
the highly deformed material undergoing fracture is affecting the transfer
of crack driving energy to the region ahead of the tip. The shear wave
speed also provides a clear indicator of the growing crack. The fracture
path is more difficult to identify in the stress plots since elements are not
removed from the simulation after the cohesive stress is reached. Stresses,
though small, remain continuous across the crack faces, making the crack
path difficult to locate. The lower plots (b) show the distribution of the
crack opening stress 22 , normalized by the cohesive stress c . These plots
show how the structure of the crack tip stress fields change as the crack
propagates through the specimen.
Figure 10 shows the crack 4.6 s after impact as it accelerates from the
arrested state to a speed of approximately 0.3 cR . In Finebergs (Fineberg
et al., 1992) results, the crack begins to display oscillatory behavior at this
speed. Neither the shear wave speed distribution (a) nor the distribution
of 22 (b) shows any signs of crack tip instability. The stress field, with a
strong stress concentration marking the current tip position, displays the
expected shape similar to the classical tip fields of Irwin (Irwin, 1957).
Figure 11 shows the crack at 5.1 s after impact, the time at which the
acoustic barrier shown in Figure 8(b) has just started to form. The initial
shape of a growing deformation-softened region is evident at the tip of the
crack. The crack speed at this instant hovers around 0.4 c R . Both the shear
CRACK DYNAMICS USING VIB METHOD 301

Figure 10. 4.6 s after impact: (a) the distribution of the instantaneous shear wave
speed normalized by the initial shear wave speed ( c s ) 0 of a wave propagating in the
X 1 -direction and (b) the distribution of the 22 component normalized by the cohesive
stress c ,both shown over the deformed domain.
302 P. A. KLEIN AND H. GAO

Figure 11. Contours of (a) the acoustical shear wave speed and (b) the opening stress
5.1 s after impact.
CRACK DYNAMICS USING VIB METHOD 303

Figure 12. Contours of (a) the acoustical shear wave speed and (b) the opening stress
6.2 s after impact.
304 P. A. KLEIN AND H. GAO

Figure 13. Contours of (a) the acoustical shear wave speed and (b) the opening stress
8.3 s after impact.
CRACK DYNAMICS USING VIB METHOD 305

wave speed and the stress distribution display waves being emitted from
the tip region in a manner that is not present in Figure 10. Figure 12 shows
the crack tip shortly after the initial branch in the fracture path. Both plots
in the figure still display a high degree of symmetry. Elastic waves in the
wake of the moving tip are even more evident. In the stress plot, the two
tips are so close to each other that the combined stress field has a shape
which resembles the field of a single tip, though the extent of the highly
stressed material is much larger. Clearly, the crack tips are interacting so
strongly that crack propagation criteria relying on classic K -field analyses
are inapplicable. The final plots in Figure 13 show the crack in a late stage
of the simulation. Several branches are clearly visible. Both plots display
a degree of chaotic behavior as each tip individually seeks a fast fracture
path. The highly stressed region ahead of the multiple crack tips extends
over a tremendous area, and no resemblance to the classical crack tip fields
remains.
Figures 14 and 15 show the results of the three-dimensional simulations.
In each figure, the upper plot (a) shows the distribution of the minimum
instantaneous shear wave speed, while the lower plot (b) shows the surface
within the domain at which the minimum shear wave speed has dropped to
half of the corresponding value in the undeformed material. Figure 14 shows
the results before the onset of the first branch. A pronounced barrier is vis-
ible ahead of the propagating crack. Although the geometry and boundary
conditions possess a degree of planar symmetry, the crack is clearly three-
dimensional in character. The regions of the crack front approaching the
free surfaces trail behind the region of the crack front nearer the center of
the domain. Figure 14 shows results shortly after branching has initiated.
The first true branches initiate on the free surfaces and propagate inward.
The three-dimensional simulations terminate before extensive branching is
observed due to difficulties with mesh tangling, elements with negative
volume produced by extreme deformations.

8. Conclusions

The simulations of dynamic crack propagation in this section show qualita-


tive agreement with experimental observations of fast fracture. The time-
averaged, terminal velocities of crack propagation are clustered around
c R / 2, though the instantaneous velocity fluctuates rapidly in the range 0.3-
0.8 c R . Initial instabilities to straight-ahead propagation appear for crack
speeds as low as 0.4 cR , speeds that cannot be predicted using classical
fracture analysis. The results provide support for the local limiting speed ex-
planation for the onset of crack tip instabilities. The VIB constitutive model
displays a continuous reduction in stiffness as the material is stretched
306 P. A. KLEIN AND H. GAO

Figure 14. The acoustical barrier in three dimensions before branching 2.7 s after
impact.

toward the cohesive limit. This elastic softening creates an unstable condi-
tion at a rapidly propagating crack tip because the straight-ahead fracture
path is blocked by an acoustic barrier to the energy transfer needed to
sustain the fracture process. In order to circumvent this barrier, the crack
seeks alternate fracture paths, leading to tip oscillations that eventually
result in larger scale branching. Notably, the local limiting speed theory
CRACK DYNAMICS USING VIB METHOD 307

Figure 15. The onset of branching in three dimensions 3.2 s after impact.

does not require the action of micro-cracks, material inhomogeneities, or


wave reflections from boundaries in order to explain the appearance of
instabilities in the fracture path. Although these effects may contribute to
fracture surface roughening in certain cases, they are not strictly required
if one adopts a hyperelastic view of the near tip deformations.
308 P. A. KLEIN AND H. GAO

References

Abraham, F.: 1997, On the transition from brittle to plastic failure in breaking a
nanocrystal under tension (NUT). Europhysics Letters 38, 103106.
Abraham, F., D. Brodbeck, R. Rafey, and W. Rudge: 1994, Instability dynamics of
fracture: a computer simulation investigation. Physical Review Letters 73, 272275.
Fineberg, J., S. Gross, M. Marder, and H. Swinney: 1991, Instability in dynamic fracture.
Physical Review Letters 67, 457460.
Fineberg, J., S. Gross, M. Marder, and H. Swinney: 1992, Instability in the propagation
of fast cracks. Physical Review B 45, 51465154.
Freund, B.: 1990, Dynamic Fracture Mechanics. New York: Cambridge University Press.
Gao, H.: 1993, Surface roughening and branching instabilities in dynamic fracture.
Journal o the Mechanics and Physics o Solids 41, 457486.
Gao, H.: 1996, A theory of local limiting speed in dynamic fracture. Journal o the
Mechanics and Physics of Solids 44, 14531474.
Gao, H.: 1997, Elastic waves in a hyperelastic solid near its plane strain equibiaxial
cohesive limit. Philosophical Magazine Letters 76, 307314.
Gao, H. and P. Klein: 1998, Numerical simulation of crack growth in an isotropic solid
with randomized internal cohesive bonds. Journal of the Mechanics and Physics of
Solids 46, 187218.
Gurson, A.: 1977, Continuum theory of ductile rupture by void nucleation and growth:
PART I. Journal of Engineering Materials and Technology 99, 215.
Hill, R.: 1962, Acceleration waves in solids. Journal of the Mechanics and Physics of
Solids 10, 116.
Irwin, G.: 1957, Analysis of stresses and strains near the end of a crack traversing a
plate. Journal of Applied Mechanics 24, 361364.
Kachanov, M.: 1992, Effective elastic properties of cracked solids: critical review of some
basic concepts. Applied Mechanics Review 45, 304.
Klein, P.: 1999, A Virtual Internal Bond Approach to Modeling Crack Nucleation and
Growth. Ph.D. thesis, Stanford University.
Klein, P. and H. Gao: 1998, Crack nucleation and growth as strain localization in a
virtual-bond continuum. Engineering Fracture Mechanics 61, 2148.
Marsden, J. E. and T. J. R. Hughes: 1983, Mathematical Foundations of Elasticity. New
York: Dover Publications, Inc.
Ravi-Chandar, K.: 1998, Dynamic fracture of nominally brittle materials. International
Journal of Fracture 90, 83102.
Ravi-Chandar, K. and W. Knauss: 1984a, An experimental investigation into dynamic
fracture: I. crack initiation and arrest. International Journal of Fracture 25, 247262.
Ravi-Chandar, K. and W. Knauss: 1984b, An experimental investigation into dynamic
fracture: II. microstructural aspects. International Journal of Fracture 26, 6580.
Ravi-Chandar, K. and W. Knauss: 1984c, An experimental investigation into dynamic
fracture: III. on steady state propagation and branching. International Journal of
Fracture 26, 141154.
Ravi-Chandar, K. and W. Knauss: 1984d, An experimental investigation into dynamic
fracture: IV. on the interaction of stress waves with propagating cracks. International
Journal of Fracture 26, 189200.
Stakgold, I.: 1950, The Cauchy relations in a molecular theory of elasticity. Quarterly
of Applied Mechanics 8, 169186.
Tadmor, E., M. Ortiz, and R. Phillips: 1996, Quasicontinuum analysis of defects in
solids. Philosophical Magazine A 73, 15291563.
CRACK DYNAMICS USING VIB METHOD 309

Xu, X.-P. and A. Needleman: 1994, Numerical simulations of fast crack growth in brittle
solids. Journal of the Mechanics and Physics of Solids 42, 13971434.
Yoffe, E.: 1951, The moving Griffith crack. Philosophical Magazine 42, 739750.
This page intentionally left blank.
CRACK TIP PLASTICITY IN COPPER SINGLE CRYSTALS

JIN YU
Department of Materials Science and Engineering
Korea Advanced Institute of Science and Technology
P.O. Box 201, Chongryang. Seoul, Korea

AND

J.W. CHO
Technical Center
Deawoo Heavy Industries Co
Inchun, Korea

Abstract: Crack tip fields in ductile crystals were studied using diffusion bonded
copper single crystals for the two orientations studied by Rice. An optical
microscope, stylus profilometer, and X-ray were used to study slip traces on
specimen surfaces, surface profiles and lattice rotations, respectively. The plastic
zone developed as an assemblage of fan-shaped sectors, the details of which
depended on the crystal orientation and the latent hardening behaviors of the crystal.
Resultantly, deformation fields of the two orientations were substantially different
from each other and also from theoretical predictions as well. Etch pit observations
of the specimen interior showed slips on the secondary (or tertiary) systems to meet
the compatibity requirement, and that crack tip plastic sectors found on specimen
surfaces are reasonably valid in the specimen interior as well, particularly for the B
orientation. A simple plane strain model based on exclusive latent hardening could
explain many features of experimental results observed on specimen surfaces and
specimen interiors reasonably.

1. Introduction

In our previous work [1], two high symmetry orientations of the f.c.c. crystals
studied by Rice [2] were investigated using diffusion bonded Cu single crystals. The
plastic zone developed as an assemblage of fan-shaped sectors for both specimens,
but observed slip traces and sector positions on specimen surfaces differed markedly
between the two orientations and also from the theoretical predictions as well [2-4].
Operations of slips on coplanar slip planes (CSP) were mutually exclusive, and
caused necking in one orientation but protrusion in the other. The disparities were
ascribed to the extensive work hardening common in f.c.c. crystals which introduced
complex problems such as latent hardening and anisotropic expansion of the yield
surface. Following experiment by Shield [5] based on the surface strain
measurement by a microscopic Moire showed identical crack tip plastic sectors
311
T.-J. Chuang and J.W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 311-329.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
312 J. YU AND J. W. CHO

found by Cho and Yu [1], and finite calculation by Mohan et al. [6] suggested that
both finite deformation and lattice rotation strongly influence the structure of the
solution. Recently, Cuitio and Ortiz [7] took into account the latent hardening
behavior of copper using a forest hardening model [8], and showed that slip
activities differed markedly between the specimen interior and the surface indicating
that the surface observations provided only indirect measures of the interior.
In the present analysis, slip trace observations on specimen surfaces were
detailed at several loading stages, and elementary observations of the crack tip
displacement fields and the lattice rotation on specimen surfaces were reported.
Then, slip traces in the specimen interior were analyzed through etch pit
observations on {111} planes using specimens with very low dislocation density. A
simple model based on the Tresca yielding and the latent hardening is proposed to
explain slip traces on the specimen surfaces and interior.

2. Experimental Procedure

High-purity Cu single crystals (99.999%) with a dislocation density of 5 1010m 2


were grown by the Bridgman method and cut into the two crystal orientations shown
in Fig. 1. In the A orientation, the crack plane is (010), the crack front lies along
-
[101], and the crack propagation is along the [101] direction. The B
-
orientation, obtained by rotating the A crystal 90 clockwise around the [101] axis,
- -
has the crack plane on the (101) and crack propagation along the [010] direction.
According to Rice [2], the yield locus is unaltered by this operation and therefore the
two orientations have identical stress and deformation fields within the small
displacement gradient formulation neglecting the lattice rotation effects. For both
- -
orientation, (111) and (111) are CSPs with a zone axis along [101] which can
- -
intersect the whole crack front line, while (1 11) and (111) are non-coplanar slip
planes (NSP) which can cross the crack front at points. For the sake of convenience,
the z=0 plane was set at the specimen middle plane as shown in Fig. 1.
A cracked bend specimen was made by cutting a single crystal into two pieces,
and joining them along the original cut planes after slight tapers were made. The
crack tip radius was typically less than 5 x 10 6 m , and tilt and rotation angle across
the bonded plane were typically less than 1. Specimens with low dislocation
density (10 9 m 2) were prepared for the etch pit observation of the specimen interior
by cyclic annealing [9], and cut along the {111} plane parallel to the crack front but
not intersecting the crack tip. Hereafter, A1/B 1 and A 2/B2 denote specimens with
high and low dislocation densities, respectively. The 3 point bend tests were
conducted under a loading rate of 1.67 106 ms 1 , and slip traces on the specimen
surface were observed with an optical microscope after unloading. Variations of the
through thickness displacement (u z ) were measured using a stylus profilometer with
a resolution of 0.1m. Displacements on specimen surfaces were measured by
recording the positions of inclusions and etch pits before and after the bend tests,
while the lattice rotations in the surface layer were measured using the back
reflection Laue method with a resolution of 0.5.
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 313

Figure 1. Two crystal orientations studied by Rice [2]; (a) A and (b) B. The { 111} planes noted by solid
lines are CSP, and those marked by dotted lines are NSP.
314 J. YU AND J.W. CHO

3. Experimental Results

3.1. Load-Displacement Curves


Single crystal tension specimens were loaded along the [010] and the [101] axes,
and resultant stress-strain curves are presented in Fig. 2(a). Both specimens showed
almost no stage I hardening, and transitions to stage II hardening occured at a tensile
stress of around ~10 MPa. If this value is taken as the yield stress( y), the plane
strain limit loads (Po) of the cracked three point bend specimens was found to be
1.12 10 4 and 1.43 104 N / m for the A and B orientations, respectively [10].
The load-displacement (P-) curves of the three point bend specimens A1 and
B 1 are shown in Fig. 2(b). Numbers in the P- curves denote serial unloading and
reloading stages, and Bauschinger effects were negligible. Note that the applied load
far exceeds the limit load even at the loading stage 1. The P- curves of the A 1 and
B1 specimens showed only slight differences because the effect of crystal anisotropy
became smaller under the triaxial stress state present near the crack tip. If the
specimens were treated as isotropic, the work hardening coefficient n (in = A n)
was deduced to be 2 from Fig. 2(b) using I1yushins theorem [10]. Here, and
refer to the uniaxial stress and strain, respectively.

3.2. Slip Traces on Specimen Surfaces

3.2.1. A 1 specimen
Evolution of the crack tip plastic zone at the four loading stages indicated in Fig.
2(b) are presented in Figs. 3(a) - (d), and surface slip traces after the fourth loading
are schematically described in Fig. 3(e). The fan shaped plastic sectors with well
defined sector boundaries developed from the load stage 1. With further load, more
slip lines appeared on the specimen surface and eventually developed into slip
bands. In general, plasticity in the sector VI was the least active except that in sector
V, and the CSP traces in the sector VI which were observed only after the loading
stage 3 were not counted here.
-
Slip traces on the (111) and (111) CSPs were confined to the sectors II and III,
- -
respectively, while those on the (111) and (111) NSPs extended over several
sectors. Operations of slips on CSPs were mutually exclusive in sectors II and III,
and slips on the (111) plane of the sector II was not observed very near the crack tip
for r 8 10 m even after the fourth loading. A close examination of the sector
4
-
boundary showed that each slip trace on the (111) plane in sector III is blocked by
another slip trace on the (111) plane in sector II, or the other way around. Spacings
between parallel slip traces in sectors II and III increased with the distance from the
crack tip, but decreased with further loading. Solid arrows in Fig. 3(e) show
- - -
dominances of the (111)[011] over the (111)[110] in sector II, and the (111)[110]
-
over the (111)[011] in sector III as necking occured, which implies that pairs of slip
- -
systems giving effective in-plane shear along the [121] or [121] directions under
plane strain deformation were not equally activated on specimen surfaces.
Slip traces on NSP developed strongly in sectors I and II from the loading stage
1; however, those in the sector III became clear only after loading stage 3. In
sectors I and IV, only NSP slips were operative, and it was not possible to
- -
distinguish the (111) and (111) NSP slip traces, which were later shown to be
mutually exclusive ( c. Fig. 8 ).
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 315

Figure 2. (a)The stress strain curves from the uniaxial tension tests with load along [010] and [101], and
(b)load-displacement curves of the 3 point bend specimens A1 and B 1 . Numbers in the curves denote
sequential unloading and reloading stages.
316 J. YU AND J. W. CHO

Figure 3. Developments of the crack tip plastic sectors of the A1 specimen after the (a) 1st, (b) 2nd, (c)
3rd, (d) 4th loading stages indicated in Figure 2. and (e) a schematic diagram of (d).
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 317

Sector boundaries maintained constant angles at the crack tip during the
4
sequence of loading, but deflected backward for r > 6 10 m ( line) due to remote
stresses. If near tip fields dominated the region with r 610 m, only CSP slip
4

traces in sectors III and NSP slip traces in sector I~III appeared to form initially by
the crack tip stress fields. Noticeable slip traces were not found in sector V, and the
elastic sectors noted by Rice [2] are a possibility.

3.2.2. B1 specimen
Here, only CSP slip traces, (111) traces in sector II and (1 1) traces in sector III
developed from loading stage 1, and the main features remained the same with
further loading. Figure 4 shows the crack tip plastic zone after the 4th stage loading
and positions of sector boundaries differ slightly from those of Shield [5],
presumably due to differences in notch radius, amount of bending, etc. The
4
boundary deflected backwards with the increase of load for r 6 10 m, and
thus only CSP slips appeared to be activated by the near tip fields (r 6 10 m ) .
4

However, subsequent etch pit analysis of the specimen interior (cf. Fig. 9.) showed
activation of slips on the NSPs to meet the compatibility requirement at sector
boundaries, and the results should be interpreted as more active NSP slips in the A
specimen than in the B specimen.
As in the A 1 specimen, slips on CSPs were mutually exclusive and expected to
operate under the plane strain condition too. It was not clear from this observation
alone whether sectors I and IV are elastic sectors noted by Rice [2], or sectors of low
plastic strain. After the 3rd loading stage, slight NSP traces on the ( 11) and (11 )
planes parallel to the y axis appeared in the sector IV. Note that mechanically
introduced scratch lines of Fig. 4(b) were deflected by 6~15 at sector boundaries
indicating strain discontinuity there. Further analysis showed that the rotation
around the z-axis based on the scratch line analysis was 3~4 times larger than the
lattice rotation found by X-ray, which suggests that the rotation by slip accounted
for most of the rotation observed here [11].
Here, based on the slip line observation on specimen surfaces which had the
stress state near the plane stress, postulates were made of those in the specimen
interior which was close to the plane strain state. Under plane strain deformation, the

effective slip vectors on the (111) and (1 1) planes can be taken as the and

and the resolved shear stress (RSS)

= b i ij n j (1)
is not affected by the in-plane stress zz because b z = n z = 0. Here, n and b are unit
vectors normal to the slip plane and along the slip direction, respectively. Thus,
variations of zz along the specimen thickness direction do not influence RSS of
CSP slips because nz equals zero, and CSP slips active under the plane stress
condition are expected to be the same under the plane strain too. In contrast, NSP
slips are expected to diminsh in the specimen interior because nz 0 and b z0 for
slips on the ( 11) and (11 ) planes, and RSS decreases with zz .
318 J. YU AND J. W. CHO

Figure 4. Crack tip plastic sector of the B1 specimen (a) after the 4th loading stage; (b) an enlarged
photograph of (a) showing scratch lines; and (c) a schematic diagram of (a).
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 319

3.3. Near Tip Surface Profiles

By running the stylus profilometer parallel to the crack line, out of plane
displacements (uz) on specimen surfaces were measured and presented in Figs. 5(a)
and (b). For the sake of convenience, the maximum necking point was set as uz = 0.

3.3.1. A1 specimen
Necking dominated the surface profiles of the near-tip regions, and was much more
severe here than in the B1 specimen. Since necking can be efficiently accommodated
by operations of slips on the NSPs, the amount of shear and corresponding RSS on
the NSPs are assumed to be greater in the A 1 specimen. The presence of the
boundary was quite clear, but the boundary was hardly observable. The maximum
necking occurred at the boundary where CSP slips intersect, and slip systems
compatible with the necking profile were (1 1)[110] in sector III and (111)[01 ] in
sector II. Therefore, in addition to the operation of NSP slips, necking is caused by
the asymmetric operations of CSP slips near specimen surfaces.

3.3.2. B1 Specimen
In contrast to the A1 specimen, slips on two adjacent CSPs caused local protrusion
here; and directions of slip in sectors II and III of the specimens A1 and B1 , which
are compatible with the surface profiles, are marked with arrows in Figs. 3 and 4.
The necking profile observed here corresponded to more frequent slip along [011]
than along [110] on the (1 1) plane {or more frequent slip along [0 1] than along
[1 0] on the (111) plane}. If only CSPs are active slip planes and the two slips
along face diagonal directions on each CSP are equally activated, the specimen
undergoes plane strain deformation and u z = 0. Thus, necking is caused primarily by
slips on NSPs and unequal activation of slips on CSPs made supplementary necking
or protrusion. Overall, the B specimen was much closer to the plane strain
deformation and thereby more suitable for the study of crack tip plasticity.

3.4. Displacement Fields Near The Crack Tip.

Using the work hardening coefficient (n) of the two specimens deduced from Fig. 2
(b), the displacement fields near the crack tip were calculated using the HRR
solution [17,18] and marked as vectors with flat ends in Fig. 6. These were
compared with experimentally measured displacements of etch pits, which were
denoted as vectors with dots and arrows at the end for A 1 and B1 specimens,
respectively. Start and end points of vectors corresponded to positions before and
after bending ( 4th stage), and both specimens were permanently bent about 3.7
after the test. Displacement vectors of the A1 specimen leaned backward (toward
higher ) compared to those of the B 1 , while calculated vectors based on the HRR
solution fell in-between. Note that vector magnitudes increased with , but that
differences between specimens diminished with (cf. encircled areas 1, and 2 in the
figure). Having different displacement fields, A1 and B 1 specimens were expected to
have different stress and strain fields and show different slip traces which were not
just the rotated product of each other.
320 J. YU AND J. W. CHO

Figure 5. Surface profiles measured by running stylus profilometer at constant y for (a) A1 and (b) B1
specimens. u z was set zero at the maximum necking point of each scan
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 321

Figure 6. Displacement vectors near the crack tip of A 1 and B 1 specimens after the 4th stage loading
denoted with solid dots and arrow tips, respectively. Calculated HRR displacements vectors denoted with
flat ends.

Figure 7. Locations on the specimen surfaces studied with X-ray for the lattice rotation measurements of
(a) A 1 and (b) B 1 specimens (30)

3.5. Lattice Rotations Near The Crack Tip

Rotations of crystal lattices after the fourth loading at various spots on the specimen
surfaces marked in Fig. 7 are summarized in Table 1

3.5.1. A 1 specimen
All the sectors except for sector I (spot ) showed clockwise rotations around the z
axis ( z ) by 3~4, which was close to the plastic bend angle during the 3 point bend
test. Rotations around the y axis ( y ) were much smaller ; y = 0.3~0.5 for the spot
and -0.5~-0.7 for the spot where the positive sign of the rotation corresponded
to the clockwise rotation. The clockwise rotation in sector I and the anticlockwise
322 J. YU AND J. W. CHO

rotation in the sector V were consistent with the observed necking behavior in Fig.
5(a). For other spots in sectors II, III, and IV, x and y were not discernable by this
method.

3.5.2. B1 specimen
As in the A1 specimen, z = 3~4 for all sectors except for the sector I (spots and
), and x and y were much smaller than z . The absence of rotation in sector I
ahead of the crack tip is consistent with the calculation by Mohan et al. [6]. For the
rotation around the x and y axes, y = 0.5~1 for the spot in the sector III, and x =
0.5~1 for the spot on the boundary, which are consistent with he necking
profile shown in Fig. 5(b). It is interesting to note that x and y decreased with
distance from the crack tip ( r ) within a sector while z was almost independent of
r. This appears partly inconsistent with the result by Mohan et al. [6] which showed
decreases of the rotation angle with r except the sector I. Presumably, z was mainly
caused by the plastic bending during the bend test, while x and y were related to
the local necking or protrusion affected by crack tip fields.

TABLE 1. Lattice rotation angle near the crack tip measured by X-ray

A 1 specimen

sector spot x y z
I 0 0.3~0.5 0

II ND ND 3~4

III ND ND 3~4

IV DN ND 3~4
V 0 -0.5~0.7 3~4

B1 specimen
sector spot x y z
I 0

II 0.5~1 0 3~4
III 0 0.5~1 3~ 4

IV 3~4

ND: not detectable

3.6. Specimen Interior Observation


The A2 and B2 specimens were cut along the {111} plane not interesting the
crack tip, and dislocation etch pits of the specimen interior were investigated. Figure
8(b) shows mosaics of dislocation etch pits of the A2 specimen on the (111) cut
plane which is 0.43 mm apart from the crack tip. The region to the left of the bonded
interface corresponded to the upper part of the crystal (i.e. y > 0, and z > 0), and vice
versa. Since and boundaries were clearly discernable and almost parallel to
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 323

Figure 8 (a) A 2 specimen cut along the (111) plane 0.43 mm apart from the crack tip; (b) dislocation etch
pits on the (111) plane for z > 0.
324 J. YU AND J. W. CHO

the bonded interface, the positions of sector boundaries and remained


almost constant regardless of the plane stress or plane strain conditions. In Figure 3,
slip traces were not found in sector II for r 0.8 mm, but a reasonably high density
of dislocation etch pits were found in sector II indicating the occurrence of slips on
the and/or NSP slip systems to meet the compatibility requirement.
The constant regardless of the plane stress or plane strain conditions. In Fig. 3,
slip boundary appeared most distinct due to the high density of sessile dislocations
produced from dislocation interactions on the two CSPs.
The figure clearly shows inclined slip traces due to the operation of NSP slips,
and slips on the left ( y > 0 and z > 0) and right side ( y < 0
and z > 0 ) of the bonded interface, respectively (see next). The NSP slips extended
over sectors (I~IV) on the specimen surface, but were very much diminished in the
specimen interior. Thus, the crack tip plastic zone was much larger on the specimen
surface than in the specimen interior because RSS on the NSP was larger for the
plane stress than the plane strain conditions. The reason why only slips on NSP
depend sensitively on the stress state is related to the fact that RSS depends on z z
for the NSP slips but not for CSP slips.
In the fatigue crack growth experiment using a Cu single crystal with the A
orientation, Neumann [14,15] found slip traces on the plane in sector III and
on the and planes in section IV on the specimen surfaces, but only
traces of sector III in the specimen interior, which is consistent with the
observation of the present work. A significant difference between the two studies is
that dislocations were generated mainly at the crack tip in Neumanns case by the
excessive work-hardening during the fatigue precracking, but at the near tip
dislocation sources in the present case.
Note that operations of NSP slips were also mutually exclusive. Among the four
NSP slip systems which can cause nonzero u z , only slip operated in the
region y > 0 and z > 0, and there was a mirror symmetry in the slip operation with
respect to the y = 0 and z = 0 planes. The selection of a NSP slip system in a given
region was dictated by the compatibility to the macroscopic deformation and the
preferential slip initiation in the highly stressed regions. Accordingly, slip
propagation from the specimen surface into the bulk and from the near tip region
into the far field region were favored, which were both along the directions of
decreasing RSS.
In the case of the B 2 specimen, a cut was made along the plane, 0.15 mm
away from the crack tip as shown in Fig. 9(a), and the resultant etch pits are shown
in Fig. 9(b). Unlike the A2 specimen, which showed active NSP slip traces near the
specimen surface, inclined slip traces coming from the slips on the NSPs were not
found, and the etch pit densities were more or less constant through the thickness.
Since sector boundaries were parallel to the bonded interface, the specimen
underwent basically the plane strain deformation by CSP slips throughout the
thickness, which was not affected by zz .
Etch pits in the sector II were much stronger than those in the sector III because
of the intersection of the (111) slip with the observation plane{ plane}, and
possible formation of Lomer -Cottrell locks. In addition, from numerous etch pits
ascribable to dislocations on the secondary slip systems, we assumed that NSP slips
also operated in sectors II and III, but to a vanishing degree. to the CSP slips.
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 325

Figure 9 (a) B 2 specimens cut along the plane 0.15 mm apart form the crack tip; (b) etch pits on the
cut plane for z>0; and (c) a magnified version of (b).
326 J. YU AND J. W. CHO

Figure 10 Crack tip plastic sectors constructed using exclusive latent hardening for the (a) A orientation
under plane stress, (b) A under plane strain, (c) B under plane stress, and (d) B under plane strain
condition.
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 327

4. Discussion

In the previous section, it was shown that the A and B specimens showed quite
different crack tip plasticity primarily due to the anisotropic expansion of the yield
locus with the latent hardening [16,17]. In order to understand the crack tip slip
behaviors of copper single crystals, a simple model based on exclusive latent
hardening is proposed here with the following assumptions:
1. Once plastic flow occurs in the primary slip system with the largest Schmid
factor, exclusive latent hardening suppress slip activities in subsequent slip systems.
2. Slips are initiated in regions of higher RSS and propagate into regions with
lower RSS, in conformity to the macroscopic plastic flow.
Using the crack tip stress fields based on anisotropic elasticity [18], contours of
the critical resolved shear stress (CRSS) were calculated for all the slip systems of
the A and B specimens under the plane stress and plane strain conditions, and the
primary slip traces were marked in Fig. 10. Note that the CRSS contours of the slip
systems producing plane strain deformation, slip systems BII and BV, DI and DVI,
AIII and CIII, are more or less the same in size under the plane stress or plane strain
conditions, while those producing non-plane strain deformation, slip system AVI
and AII, CI and CV, are much larger under the plane stress condition. Note also that
the non-plane strain CRSS contours are much larger in the A specimen, which
explains why NSP shear and the degree of necking are much larger in the A
specimen.
In the case of the A specimen, the plane stress prediction is quite different from
Fig. 3 except extensive NSP slips traces, while the plane strain prediction can
reasonably describe CSP slip traces observed in the sectors II and III, and NSP slip
traces in sector I and IV. Needless to say, coexistence of NSP and CSP slips in
sectors II and III could not be predicted due to the exclusive latent hardening
assumption.
In the case of the B specimen, plane stress and plane strain predictions are not
much different, and agreements are generally much better. CSP slip traces in sectors
II and III are common to the plane stress and plane strain predictions, which partly
explain the parallel , and lines throughout the specimen thickness in Fig. 9.
Overall, the plane strain prediction was closer to what was found in Fig. 4. The
reason why the plane strain predictions make better estimates of surface slip traces
observed for both specimens can be related to the necking which introduces non-
zero x z , y z , and z z, thereby deviating the stress state from the plane stress state
substantially. According to Cuitio and Ortiz [7], stress states on the specimen
surface and interior middle plane differ markedly from the plane strain field.

5. Conclusions

1. The crack tip plasticity developed in the fan-shaped sectors with well defined
sector boundaries, but the two orientations studied by Rice[2] showed quite different
deformation fields; The B specimen showed only CSP sectors, while A specimen
showed CSP and NSP sectors. Operations of slips on CSPs were mutually exclusive
and the same was true of NSPs, even though CSP and NSP slips operated
simultaneously in sectors II and III of the A specimen. NSP slips were quite active
on the A specimen surface due to the large RSS and necking. Also, operations of
slips on CSPs caused local necking in the A specimen but protrusion in B.
Displacements were continuous at sector boundaries but not the displacement
328 J. YU AND J. W. CHO

gradient, which suggests constant plastic strain within a sector but strain
discontinuity at sector boundaries.
2. Etch pit observations of near tip displacement on specimen surfaces
again confirmed that the two orientations have quite different crack tip fields.
Generally, displacement vectors of the A specimen pointed toward higher angle and
differences between the two specimen diminished with . Subsequent X-ray
measurement showed that both specimens had z = 0 in sector I but z = 3~4 in all
other sectors which was close to the permanent bend angle after the test, suggesting
that most of the rotation near the crack tip was caused by slip. Rotation of the lattice
due to necking was typically smaller than 1.
3. Etch pit observations of the specimen interior showed that crack tip
sectors found on specimen surfaces were reasonably valid in the specimen interior
as well, particularly for the B specimen. In the case of A specimen, NSP slips
developed near the surface but diminished in the specimen center with decreasing
z z . Sectors showing only single slip traces on the specimen surface, for example
sector II of the specimen B, revealed secondary slip traces attesting the limitations
of the experimental method used here.
4. A plane strain model based on exclusive latent hardening could explain
experimental observations of the primary slip traces and sector boundaries on
specimen surfaces and interior reasonably well.

6. References

1. Cho, J.W. and Yu, J.: Near crack tip deformation in copper single crystals, Phil.
Mag. Lett. 64 (1991), 175-182
2. Rice, J.R.: Tensile crack tip fields in elastic-ideally plastic crystals, Mechanics of
Materials 6 (1987) 317-335
3. Saeedvafa, M. and Rice, J.R.: Crack tip singular fields in ductile crystals with
taylor power-law hardening .2. plane-strain, J. Mech. Phys. Solids 37 (1989),
673-691
4. Rice, J.R., Hawk, D.E. and Asaro, R.J.: Crack tip fields in ductile crystals, Int.
J.Fracture 42 (1990), 301-321
5. Shield, T.W.: An Experimental study of the plastic strain fields near a notch tip in
a copper single crystal during loading, Acta Mater. 44 (1996), 1547-1561
6. Mohan, R.,Ortiz, M. and Shih, C.F.: An analysis of cracks in ductile single-
crystals .2. mode-I loading, J. Mech. Phys. Solids 40 (1992) 315-337
7. Cuiti o, A.M. and Ortiz, M.: Three-dimensional crack-tip fields in four-point-
bending copper single-crystal specimens, J. Mech. Phys. Solids 44 (1996) 863-904
8. Cuitio, A.M. and Ortiz, M.: Computaional modeling of single-crystals, Modeling
Simul. Mat. Sci. Eng. 1 (1993) 225-263
9. Kitajima, S., Ohta, M. and Tonda, M.: Production of highly perfect copper
crystals with thermal cyclic annealing, J. Cryst. Growth 24/25 (1974) 521-526
10. Kanninen, M.F.: Advanced Fracture Mechanics, Oxford University Press., N.Y.,
1985
11. Yu, Jin, unpublished work, 1991
12. Rice, J.R. and Rosengren, G.F.: Plane strain deformation near a crack tip in a
power-law hardening material, J.Mech. Phys. Solids. 16 (1968) 1-12
13. Hutchinson, J.W.: Singular behaviour at the end of a tensile crack in a hardening
material, J. Mech. Phys. Solid 16 (1968) 13-31
CRACK TIP PLASTICITY IN Cu SINGLE CRYSTALS 329

14. Vehoff, H. and Neumann, P.: In situ sem experiments concerning the mechanism
of ductile crack growth, Acta Metallurgica, 27, (1979) 915-920
15. Neumann, P., Fuhirott, H. and Vehoff, H.: Experiments concerning brittle,
ductile, and environmentally controlled fatigue crack growth, in J.T. Fong (ed.),
Fatigue Mechanisms, ASTM STP 675, (1979) 371-395
16. Jackson, P.J. and Basinski, Z.S. : Latent hardening and the flow stress in copper
single crystals, Can. J. Phys. 45, (1967) 707
17. Basinski, S.J. and Basinski Z.S.: Chapter 16, Plastic deformation and work
hardening, P. 261 in Dislocations in Solids, Vol. 4, ed. F.R.N. Nabarro, North-
Holland, 1983
18. Paris, P.C. and Sih, G.C.: Stress analysis of cracks, in Fracture Toughness
Testing, ASTM STP 381, (1965) 30-83
This page intentionally left blank.
NUMERICAL SIMULATIONS OF SUBCRITICAL CRACK
GROWTH BY STRESS CORROSION IN AN ELASTIC SOLID

Z. TANG AND A.F. BOWER


Division of Engineering
Brown University
Providence, RI 02912
AND
T.-J. CHUANG
Ceramics Division
National Institute of Standards and Technology
Gaithersburg, MD 20899-8521

Abstract:
A front-tracking finite element method is used to compute the evolution of a
crack-like defect that propagates by stress driven corrosion in an isotropic, linear
elastic solid. Depending on material properties, loading, and temperature, we
observe three possible behaviors for the flaw: (i) gross blunting at the crack tip;
(ii) stable, quasi-steady state notch-like growth; and (iii) unstable sharpening of
the crack tip. The range of material parameters and loadings that cause each type
of behavior is computed. Our results also confirm the existence of a threshold stress
level (known as the fatigue limit) that leads to crack sharpening and ultimately
to catastrophic fracture. Contrary to earlier predictions, however, our simulations
show that the fatigue threshold is determined not only by the driving force for
crack extension but also by the kinetics associated with the chemical reaction at
the crack tip. Our results suggest that the fatigue threshold is likely to decrease
as temperature is reduced. Finally, we have computed the steady state crack tip
velocity as a function of applied load in the regime of steady state crack growth.
Our predicted crack growth law is in good qualitative agreement with experiment,
but uncertainties in material data make quantitative comparison difficult.

1 . Introduction
Advanced ceramics, fiber reinforced composites and optical glasses are exploited in
the design of devices and components by various industries, ranging from aerospace
applications to computer hardware. The durability of ceramics and glasses in ser-
vice is therefore a major concern. Experiments suggest that the lifetimes of many
components are limited by subcritical crack growth (Zhou and Curtin, 1995). Un-
der sustained loading conditions, two classes of crack growth are observed, depend-
ing on the stress, temperature and material. At elevated temperatures, the most
common form of failure is by crack growth along interfaces or grain boundaries. In
contrast, at room temperature, or in a corrosive environment, transgranular frac-
ture is the dominant mechanism of failure. In amorphous materials such as glass,
331
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 331348.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
332 Z. TANG, A. BOWER AND T.-J. CHUANG

subcritical crack growth is the main mode of failure at all temperatures. The focus
of this paper is subcritical transgranular crack growth in brittle elastic solids.
In general, ceramics and glasses are notable for their resistance to a hostile
environment. Nevertheless, if a stressed ceramic component is exposed to chemical
attack, it may suffer from a form of delayed fracture known as static fatigue. In
some materials this behavior has been attributed to a process involving chemical
dissolution of material from the region near the tips of small pre-existing cracks
in the solid (Wiederhorn, 1975, White, et al., 1986, Simmons and Freiman, 1986,
Gehrke, et al., 1990,). In this case, the loss of material causes cracks to progres-
sively sharpen and increase in length until catastrophic fracture occurs. Experi-
ments have revealed that, for these materials and ambient environment, the time
to failure is a strong function of the applied stress. In particular, if the applied
stress lies below a threshold value, known as the fatigue limit or stress corrosion
limit, failure can be avoided. It is clearly desirable to determine this limit, and in
situations where the stress must exceed the corrosion limit, to determine the rate
of crack growth as a function of the applied stress. In the literature, the latter is
often expressed empirically as v = AK n or v = B exp( CK) from the experimental
data (see for example, Wiederhorn et al. 1974, Freiman et al. 1985). The main
objective of the present work is to present a physics-based model to describe the
subcritical crack growth behavior within a material subject to chemical dissolution.
Charles and Hillig (1962) were the first to develop a micromechanical model
of crack growth by corrosion. They considered the behavior of an elliptical cavity
in an isotropic, linear elastic solid, using a model based on absolute reaction-rate
theory to characterize the rate of dissolution of material from the crack flanks
(Hillig and Charles 1965). The essential feature of this model is that the rate of
material loss from a surface is influenced by both the stress acting tangent to the
surface and also by surface curvature. Stress generally tends to increase the rate of
material loss, while the curvature of a concave surface reduces it. The competition
between these two effects may be characterized by a dimensionless parameter

(1)
where tip is some measure of the stress near the cavity tip, is the free energy
per unit area of the unstressed surface, and tip is the surface curvature near the
crack tip. For large , the effects of stress dominate over curvature, so that the
ellipse tip propagates more rapidly than the flanks. This causes the ellipse to
sharpen, and eventually results in the formation of a crack which triggers brittle
fracture. For small , the effects of stress are negligible. Material at the tip of
the ellipse then dissolves more slowly than material near the flanks, and the ellipse
is blunted, eventually evolving to a rounded cavity. The critical value of that
discriminates between blunting and sharpening gives the static fatigue limit for
the solid. Charles and Hillig estimated the critical by assuming that the crack
remains elliptical throughout its evolution.
Chuang and Fuller (1992) extended the Charles-Hillig (1962) model to com-
pute the initial rate of dissolution of material from the entire surface of the ellipse.
STRESS-ASSISTED CORROSIVE CRACK GROWTH 333

They showed that an initially elliptical flaw is unlikely to remain elliptical through-
out its growth, and instead predicted four possible regimes of behavior for the crack:
(a) gross blunting, where the rate of dissolution of material from the crack flanks
exceeds the rate of removal near the tip, so that the ellipse approaches a circular
shape; (b) enhanced blunting, where the material just adjacent to the crack tip is
removed faster than material at the crack tip itself, resulting in blunting near the
apex; (c) necking, where the material removal rate is a minimum just adjacent to
the crack tip, resulting in a neck-like crack forming near the apex and (d) gross
sharpening, where material is removed most rapidly near the crack tip. In addition
they showed that, for a typical reaction theory based consititutive law of corrosion,
a second material parameter m plays an important role in governing the behavior
of the crack. This parameter will be defined and discussed in more detail in Sec-
tion 4: for now it is sufficient to note that m quantifies the nature of the corrosion
law. In general, the expressions for both the driving force for material removal,
and also the associated activation energy, contain linear and quadratic terms in
stress. For large m, the linear term dominates, while for small m the quadratic
term is dominant. Chuang and Fullers (1992) computations suggest that there
exists a threshold value for m, which controls a transition from enhanced blunting
behavior (regime b) to neck-like crack growth (regime c).
Existing micromechanical models are thus based either on a simple geometrical
description of the crack, or draw conclusions based on the initial rate of material
loss from the crack surface. In this paper, we use a numerical technique to compute
in detail the evolution of a crack propagating by stress driven corrosion. We
consider a large, plane, linear elastic solid which contains a crack-like notch near
its center. We adopt Hillig and Charles (1965) constitutive law to describe the
rate of material loss from the crack surface as a function of stress and curvature.
The finite element method is used to solve the coupled equations of linear elasticity
and those governing material loss from the crack surface, while a front tracking
method is devised to track the evolution of the cracks geometry with time.
Our results confirm many of the predictions of existing models: we observe
a transition from crack blunting to sharpening at a critical value of ; we find
that m has a strong influence on the behavior of the crack, and observe most of
the features of crack evolution predicted by Chuang and Fuller (1992). However,
some surprising new insights emerge from our simulations. In particular, we find
that the transition from blunting to sharpening is determined not only by the
driving force for material removal, but also by the kinetics of this process, so that
there is no single pair of values for and m which lead to crack sharpening.
Instead, the critical combination of and m depends on a third dimensionless
material parameter , which is a function of temperature. The implication of
this result is that the fatigue threshold for a given material is likely to vary with
temperature: our results suggest that a decrease in temperature will decrease
the fatigue threshold. Secondly, our results show that the initial rate of material
loss from the crack surface is not a good predictor of its subsequent behavior.
Consequently, the four regimes of behavior proposed by Chuang and Fuller (1992)
334 Z. TANG, A. BOWER AND T.-J. CHUANG

Fig. 1. Idealized geometry used to study the growth of a stress corrosion crack
in an elastic solid, showing a typical finite element mesh. The crack is
elliptical, with ratio of semiaxes b / a = 0.01. Each element shown is a six
noded triangle.

are observed only for a vanishingly short time. Enhanced blunting (regime b)
quickly evolves to gross blunting (regime a); and gross sharpening (regime d) is
never observed - crack sharpening is always accompanied by the formation of a neck
near the crack tip (regime c). Instead, we observe three types of crack growth:
(i) Gross blunting; (ii) Stable, quasi-steady state notch like crack growth; and
(iii) Sharpening, accompanied by the formation of a neck near the crack tip. We
find that regime (i) will occur in all materials, provided that the applied load is
sufficiently low. In contrast, regime (ii) exists only in materials in which m, exceeds
a critical threshold. In such materials, the crack will blunt at low loads, or sharpen
to propagate as a self-similar notch at higher loads. The crack would presumably
continue to grow in this manner until the stress near the crack tip exceeds the ideal
strength of the solid. In materials with m below the critical threshold, the ellipse
appears to sharpen without limit, and rapidly forms an ideal crack.

2 . Model Description
We idealize a typical ceramic component as a planar, isotropic, linear elastic solid
with Youngs Modulus E and Poissons ratio v, Fig 1. The solid is assumed to
contain a single crack like flaw, with characteristic length 2a, near its center. In
this paper, we will report results only for an initially elliptical cavity, but we have
obtained similar results for a notch-like flaw with constant tip curvature and flat
sides. The solid is loaded by a uniform remote stress acting perpendicular to
the major axis of the flaw, thereby inducing a displacement field ui (x j ) and stress
distribution i j (x j ) within the solid. The displacement and stress fields are related
STRESS-ASSISTED CORROSIVE CRACK GROWTH. 335

by the usual linear elastic constitutive law

(2)

and the stress must satisfy the equilibrium equations ij , j = 0. In subsequent


discussions, we will assume that the displacements u i are small, implying that
the change in shape of the cavity surface due to elastic distorsion of the solid is
negligible. We will, however, account rigorously for large changes in shape of the
reference configuration as material is dissolved near the crack tip.
The defect surface is assumed to be exposed to an unspecified, chemically
reactive species, which progressively removes material from the solid. We use
Hillig and Charles (1965) constitutive law to characterize the resulting rate of loss
of material. In developing this model, it is assumed that chemical reactions at the
solid/vapor interface limit the rate of material removal, so that it is not necessary
to account explicitly for processes involving diffusion of material to or away from
the reaction site, nor is it necessary to model adsorption or desorption of chemically
reactive species at the surface. The reaction is driven by a difference in chemical
potential between the material near the solid surface and the reaction product, and
the rate of reaction is determined by a combination of the driving force and the
activation energy associated with the chemical process. In the absence of stress,
this causes a flat surface to recede at uniform rate v0 , which is generally a function
of temperature. Surface curvature and stress modify the chemical potential of
atoms near the void surface, and also influence the activation energy. The recession
rate of a curved, stressed surface is therefore expressed as

(3)
Here, v n is the normal velocity of the surface in the unstressed reference configu-
ration (v n is negative because the surface is receding), R is the gas constant and
T is temperature. In addition,

(4)
where = i j ti t j is the tangential stress at the solid surface; V m is the molar
volume of the solid; is a dimensionless phenomenological constant such that
V m = ( / ) =0 is the activation volume for the reaction; is a second di-
mensionless constant, which accounts for both a quadratic term in stress in the
Taylor expansion of the activation energy about = 0 and also for the strain
energy released as atoms are removed from the surface; E' = E /(1 v 2 ) is the
plane strain modulus; is the energy per unit area of a stress free surface and
is the surface curvature, defined so that > 0 for a concave surface. Additional
details concerning the derivation of this kinetic law may be found in Hillig and
Charles (1965).

3 . Numerical procedure
336 Z. TANG, A. BOWER AND T.-J. CHUANG

We have used the finite element method to solve the equations outlined in the pre-
ceding section. It is convenient to divide the calculation into three steps. Assume
that the shape of the crack or notch at time t = 0 is known. The first step is
then to determine the distribution of stress in the solid at time t = 0. Next, we
determine the material lost from the void surface during a subsequent interval of
time t. Finally, the reference configuration is updated, and a new distribution of
stress is computed for time t = t. The computation is repeated to determine the
evolution of the crack as a function of time.
The standard finite element method for linear elastic solids is used to compute
the distribution of stress in the solid. We also use a finite element procedure
to calculate the change in shape of the void surface as material is removed by
corrosion. Let h(s) denote the depth of material lost during a time interval t
at position s on the void surface. From the preceding section, we have that

(5)

Due to the presence of surface curvature in the expression for , it is difficult to


integrate this equation with respect to time using an explicit Euler scheme. We
therefore use a semi-implicit method, noting that the change in curvature of the
surface during a time interval t may be estimated as

(6)

We use this estimate in a general Euler time integration scheme

(7)

where 0 1 is a parameter controlling the time integration. For = 0,


(7) reduces to a standard explicit forward-Euler scheme, while choosing = 1
corresponds to a semi-implicit scheme with a first order predictor for curvature.
Choosing > 0 has a marked impact on the stability of the algorithm, allowing
time step sizes to be increased by several orders of magnitude without loss of
accuracy. Our tests show that = 1 leads to the best numerical stability, while
the best accuracy appears to occur around = 0.5. The accuracy is relatively
insensitive to , however, since small time steps must be taken to ensure that the
stress field is updated correctly. We have used = 1 in all the computations
reported here.
Combining (6) and (7) and writing the result in weak form then leads to

(8)
STRESS-ASSISTED CORROSIVE CRACK GROWTH 337

Here, h(s) is a twice continuously differentiable test function of arc length around
the void surface, and S denotes the void surface. Eq (8) must be satisfied for all
admissible h(s). The usual finite element procedure is used to obtain a discrete
form of (8): the variations of h(s) and h( s ) are interpolated between discrete
points on the void surface by means of piecewise cubic Hermitian interpolation
functions, which allows the surface integrals to be expressed in terms of a finite set
of values of h, h and h/s, h/s. The condition that (8) must hold for
all h then leads to a sparse, unsymmetric, system of linear equations to be solved
for the discrete values of h and h/s.
These results then form the basis for a finite element solution for the shape of
the corrosion crack as a function of time. At time t = 0, the initial shape of the
crack is specified by a set of control points on the void surface. The geometry of
the solid is then interpolated between these points, using cubic parametric splines.
The analysis begins by generating a mesh of six noded, triangular finite elements
to fill the solid. We have found that the advancing front method of Peraire et
al (1987) is particularly effective for this purpose. The algorithm allows one to
generate meshes with an arbitrary variation of element size: in our computations
we use the error estimate of Zhu and Zinkiewicz and Zhu (1987) to generate a
nearly optimal mesh at each time step. A typical finite element mesh is illustrated
in Fig.1: the mesh contains approximately 3500 elements and 15000 degrees of
freedom. The smallest element near the crack tip has a height of approximately
6
10 a, where a is the crack length. For a crack length of 10m, this corresponds
to a spacing between nodes of only 0.01nm.
We then proceed to compute a finite element estimate for the nodal values of
displacement, and subsequently use a variational recovery scheme to project values
of stress from the integration points within each element to the nodes. The nodes
that lie on the void surface are then used to generate a one-dimesional finite element
mesh to solve (8). Finally, nodal values of h and h/s on this mesh are used
to compute a new spline representation for the void surface. The procedure is
repeated to determine the history of crack propagation.

4. Results and Discussion


To discriminate between the various regimes of behavior of the crack, we adopt the
following dimensionless measures of stress, material properties and crack geometry

(9)

Here, is the crack tip stress intensity factor, 0 is the initial crack
tip curvature, is the surface energy, E' is the plane strain modulus, V m is the
molar volume of the solid, R is the gas constant and T is temperature. Finally,
and are the two dimensionless parameters appearing in the corrosion law (34).
Note that one may also define an additional (but not independent) dimensionless
parameter
(10)
338 Z. TANG, A. BOWER AND T.-J. CHUANG
TABLE 1: Representative values for material properties

which is a function only of the material properties and is independent of applied


loading. In addition, we introduce the dimensionless time measure

(11)

Provided that the conditions necessary for the applicability of linear elastic
fracture mechanics are met, the values of and , together with any two of the
parameters , , or m, completely characterize the behavior of the crack. It is
straightforward to appreciate their physical significance: quantifies the relative
effects of crack tip curvature and crack tip stress on the rate of material removal
by corrosion: for large , stress dominates, tending to cause rapid crack growth,
while for small crack growth is retarded by the influence of curvature. Similarly,
can be loosely thought of as the ratio of crack tip energy release rate
to the Griffiths toughness Large positive values of imply a large driving
force for crack growth. However, because the chemical reaction at the crack tip
may either provide an additional thermodynamic driving force for crack growth,
or may involve additional energy dissipation as heat, the condition need
not be satisfied for the crack to advance, and crack growth is possible for all values
of . Indeed, since the kinetic parameter may be negative, for some materials
it is possible that < 0. The dimensionless parameter describes the kinetics
of crack growth: a large value for implies a large change in crack tip velocity
with stress or curvature. Finally, the material parameter m quantifies the relative
STRESS-ASSISTED CORROSIVE CRACK GROWTH 339

Fig. 2. Sequences of crack surface profiles for = 0.24, m = , = 10 4 for


two load levels: (a) Blunting occurs for = 0.8; the time interval t  =
2 10 4 . (b) Sharpening, followed by stable notch like growth occurs for
= 4.0; t = 4.8 10 5 .

magnitudes of the linear and quadratic terms in the driving force for stress driven
material removal. For m < 0, the linear term tends to increase the rate of crack
growth with stress, while the quadratic term retards growth. For m = 0, the linear
term has no contribution, reducing the corrosion law to the form used by Wilkins
and Dutton (1976). For m > 0, both linear and quadratic terms in the expansion
tend increase the rate of crack growth, and in the limit m the linear term in
the corrosion law dominates.
Typical values for the material parameters in our model are listed in Table 1.
We now proceed to investigate the behavior of the crack for the range of physically
reasonable values of the dimensionless parameters. For simplicity, we will consider
first the case = 0 ( m ), wherein the linear term in stress in the expression
for the driving force for crack growth dominates over the quadratic term. Fig. 2
shows the behavior of the crack for two different levels of applied stress , and for
an intermediate value of the kinetic parameter . The figures each show a sequence
of profiles of the crack surface (in the undeformed configuration), at equally spaced
intervals of dimensionless time t . For low values of , the crack blunts; while if
exceeds a critical threshold, the crack sharpens, forming a neck in the process.
Further insight into the behavior of the crack can be gained by examining the
distribution of surface curvature and the rate of loss of material from the region
near the crack tip. Fig. 3 shows a sequence of graphs of normalized surface velocity
as a function of arc length near the crack tip ( s = 0 corresponds to the crack tip;
the arc length is normalized by initial crack length a ); Fig.4 shows a corresponding
sequence of surface curvature. Observe that in Fig 3a, the initial velocity of the
crack tip is less than the velocity of points just adjacent to the tip. This simulation
therefore lies in the regime classified as enhanced blunting by Chuang and Fuller:
the tip initially propagates more slowly than the crack flanks. Our results show,
340 Z. TANG, A. BOWER AND T.-J. CHUANG

Fig. 3. Sequences of normalized crack surface velocity for = 0.24, m = ,


= 10 4 at two load levels: (a) Blunting for = 0.8; time interval t =
2 10 4 ; (b) Sharpening to steady notch like growth for = 4.0; t =
4.8 10 5

Fig. 4. Sequences of crack surface curvature for = 0.24, m = , = 10 4 at two


load levels: (a) Blunting for = 0.8; time interval t = 2 10 4 ; (b)
Sharpening to steady notch like growth for = 4.0; t  = 4.8 10 5 .

however, that this blunting behavior lasts for only a short time, and soon gives
way to gross blunting, where the entire tip region propagates more slowly than the
crack flanks. This is accompanied by a progressive decrease in crack tip curvature,
Fig 4a.
Figs 2b, 3b and 4b show the behavior of the crack for a high value of .
Observe that in these results, the initial velocity of the crack surface is a maximum
at the crack tip - thus placing the simulation in Chuang and Fullers (1992) regime
(d): gross sharpening. In fact, we have not observed gross sharpening in any of our
STRESS-ASSISTED CORROSIVE CRACK GROWTH 341

Fig. 5 Variation of (a) crack tip velocity and (b) curvature, for various material
parameters and load levels.

simulations: instead crack sharpening is always accompanied by the formation of a


neck near the crack tip, as shown in Fig. 2b. Figs 3b and 4b show the corresponding
variation of crack surface velocity and curvature. The crack tip velocity, curvature
and stress all increase, but the crack tip curvature increases more rapidly than
the stress, so that after a transient period the curvature and velocity distributions
approach a quasi-steady state. Further evidence for this behavior is presented in
Fig. 5, which shows the crack tip velocity and curvature for various combinations
of , m and . For = 4, = 0.24, and m = , the crack tip curvature and
velocity appear to approach steady values. We take this as an indication of stable,
quasi-steady notch like crack growth, which will continue until stress levels near the
crack tip approach the ideal strength of the solid and so trigger unstable fracture.
This form of crack growth should be contrasted with unstable sharpening, wherein
the crack tip curvature and velocity increase without limit, leading to rapid failure.
Since we do not observe either enhanced blunting or gross sharpening in our
simulations, we will not follow Chuang and Fullers (1992) characterization of the
behavior of the crack. Instead, we will classify the behavior of the crack in one of
three regimes: (i) Gross blunting; (ii) stable, quasi-steady notch like crack growth;
(iii) unstable sharpening, wherein the crack tip curvarure and velocity both increase
without limit.
We turn next to examine the influence of kinetics on crack growth behavior.
Fig 6 shows the behavior of the crack for identical values of remote stress and
material parameter m = 0, but two different values of . Since influences the
relative magnitudes of the cracks surface velocity at its tip and flanks, in this case
increasing has a qualitatively similar effect to increasing . For low values of
, the crack always blunts; while for high values of the crack tip sharpens. An
important consequence of this observation is that the critical value of required
to trigger crack sharpening depends on . Since is temperature dependent, our
simulations suggest that the fatigue threshold will vary with temperature. Lower
342 Z. TANG, A. BOWER AND T.-J. CHUANG

Fig. 6. Sequences of crack surface profiles for = 2.66, m = , = 10 4 for


increasing rate parameters (a) Blunting for = 0.24; time interval t  =
2.2 10 4 (b) Sharpening for = 0.48; t  = 5.0 105 .

Fig. 7. A fracture mechanism map showing the range of values of and required
to cause crack blunting or sharpening, for m = . In this case a sharpening
crack stabilizes to steady notch like growth.

temperatures increase and therefore decrease the fatigue threshold. Of course,


the rate of crack growth is also reduced if temperature is reduced, so that unstable
fracture will be kinetically limited at very low temperatures.
We have conducted several simulations to map the critical combinations of
and that will cause the crack to sharpen. The result is shown in Fig. 7.
Our computations suggest that there is a critical load level below which crack
STRESS-ASSISTED CORROSIVE CRACK GROWTH 343

Fig. 8. Sequences of normalized crack surface velocity for m = 0, = 10 4 at two


load levels: (a) Blunting for = 4.55, = 0.12; time interval t  = 10 4
(b) Unstable sharpening for = 4.55, = 0.24; t  = 7 10 6 .

sharpening will not occur for any . This critical stress appears to coincide with
Chuang and Fullers estimate for the fatigue threshold c r , shown as a dashed line
in Fig. 7.
We have also investigated the role of the material parameter m in governing
crack growth behavior. As an example, we next present results for m, = 0, wherein
the quadratic term in stress in (4) dominates over the linear term. For this case
the load level must be parameterized by the dimensionless group , since = 0
for all stress levels. Typical results for two values of and an intermediate value of
are presented in Fig. 8. Qualitatively, the behavior of the crack is similar to the
results presented for m = . For low loads, the crack blunts, while for high loads,
the crack sharpens. However, in this case a sharpening crack never stabilizes:
instead, the crack tip curvature and velocity both increase to the limit of the
resolution of our finite element method. Evidence for this assertion is presented
in Fig. 5, which shows variations of crack tip curvature and velocity for various
combinations of material parameters and applied load levels. For m = 0, both
the crack tip curvature and velocity appear to increase without limit if the load
exceeds the fatigue threshold.
Our estimate for the critical combinations of and which lead to unstable
crack sharpening are shown in Fig. 9. As for m = , we note that the fatigue
threshold is generally a function of , and consequently is a function of tempera-
ture. There appears to be a critical value for below which blunting always occurs,
irrespective of the value of . However, in this case the threshold does not appear
to coincide with Chuang and Fullers (1992) estimate of the fatigue threshold.
We have conducted several further simulations to investigate crack growth
behavior for arbitrary m values. Our results are summarized on Fig. 10, which
shows fracture mechanism maps for several m values. As before, parameterizes
the magnitude of the applied load, while is primarily dependent on material
344 Z. TANG, A. BOWER AND T.-J. CHUANG

Fig. 9. A fracture mechanism map showing the range of values of and required
to cause crack blunting or sharpening, for m = 0. In this case sharpening
continues without limit to form an ideal crack.

Fig. 10. A fracture mechanism map showing the range of values of and required
to cause crack blunting or sharpening, for various m. For m exceeding
between 2.5 and 3.3, a sharpening crack stabilizes; for m below this range
sharpening is unstable.

properties associated with the rate of corrosion. Recall that setting m = 0 gives
a corrosion law with only quadratic terms in stress; while for m , the linear
STRESS-ASSISTED CORROSIVE CRACK GROWTH 345

Fig. 11. Transient variation of crack tip velocity during convergence to steady-state
notch like growth, for two initial conditions. Both results have = 37.8
and so eventually converge to the same velocity.

term dominates. We find that reducing m tends to reduce the critical and that
will ensure crack blunting. For high values of m, exceeding these values will cause
the crack to sharpen, but the sharpening will stabilize to produce stable notch-like
crack growth at constant velocity and crack tip curvature. If m falls below a value
of between 2.5 and 3.3, then we see no tendency for the crack to stabilize. Instead,
the crack continues to sharpen, with a corresponding increase in crack tip velocity,
until our simulations can no longer reliably resolve the crack tip.
In the regime of stable notch-like crack growth, the variation of crack tip
velocity with applied load is of particular interest. Dimensional considerations
indicate that the steady state crack tip velocity and curvature may be expressed
as
(12)

where G and F are functions to be determined, s s denotes the steady-state crack


tip curvature, and

(13)

Fig. 11 illustrates this trend. The figure shows the variation of crack tip velocity
with time, for two cracks with identical values of crack driving force = 37.8,
but different initial conditions ( = 2.48, = 6.1) and ( = 4.98, = 1.5),
respectively. In both cases m = M = and = 10 4 . After an short transient
period, both cracks propagate with the same crack tip velocity, regardless of the
initial conditions.
346 Z. TANG, A. BOWER AND T.-J. CHUANG

Fig. 12.(a) Steady state crack tip velocity and (b) curvature as a function of applied
load, for m = M = . The scatter in the numerical data is caused by small
fluctuations in the numerical solution due to variations in finite element
mesh size.

We have calculated the steady state crack tip velocity and curvature as func-
tions of load parameter , for the particular case m = M = . Results are shown
in Fig. 12, and suggest that our numerical data may be approximated by a crack
tip velocity law of the form

(14)

This is in remarkable qualitative agreement with experimental observations, which


are generally fit by v = B e x p (CK I ), where B and C are empirical constants
that depend on temperature and nature of the corrosive environment (Wiederhorn
et al 1974, Wiederhorn 1975). Quantitative agreement is less satisfactory, however.
Experiments indicate that B ~ 10 21 ms 1 and (Wiederhorn
1
et al, 1974), while data listed in Table 1 suggest that B ~ 10 12 ms a n d C ~
This discrepancy may be partly due to errors in values for material
properties listed in the table: our predictions are particularly sensitive to variations
in and . There are several other explanations, however. Using data in Table 1,
our calculations predict steady state crack tip curvatures of the order 1011 m 1 , and
the crack tip stress is correspondingly high. The validity of a continuum linear
elastic solution in this regime is questionable. In addition, our stress-corrosion
law is somewhat speculative, and requires experimental verification. These are
promising areas for future study.

7. Conclusions
We have used a numerical technique to predict the evolution in shape of a crack
like defect propagating by stress driven corrosion. Our computations predict three
possible types of behavior for the crack: (i) gross blunting, where the crack evolves
STRESS-ASSISTED CORROSIVE CRACK GROWTH 347

towards a rounded profile, (e.g. Fig. 3a); (ii) stable, notch like crack growth,
where the crack initially sharpens, but approaches a steady self-similar profile and
propagates with constant tip curvature and tip velocity (e.g. Fig. 3b); and (iii)
unstable sharpening, where the crack tip curvature, stress, and velocity appear
to increase without limit (e.g. Fig. 8b). In general, the behavior of the crack is
determined by a dimensionless load factor , two material parameters m and ,
and a shape factor , defined in (9). The various regimes of behavior are plotted
as a function of these parameters in Figs 7, 9 and 10. For sufficiently low values of
applied load, the crack always blunts, irrespective of the value of or m. For larger
applied loads, the flaw will either sharpen to form a stable notch that propagates
with constant velocity and tip curvature, or else will sharpen without limit to
form an ideal crack. The former behavior occurs in materials with m exceeding
a threshold value between m crit = 2.5 and 3.3; in materials with m < m crit , the
crack sharpens in an unstable manner.
Our simulations confirm the existence of a critical level of applied stress which
must be exceeded to cause crack growth. Contrary to earlier predictions, however,
our results suggest that the fatigue threshold is determined not only by the driving
force for crack extension but also by the kinetics associated with the chemical
reaction at the crack tip. The critical values of or required to cause sharpening
are thus a function of m and , as illustrated in Figs 710. Since the critical stress
depends on , which is in turn temperature dependent, our results imply that the
fatigue threshold decreases as temperature is reduced.
Finally, we have used our computations to calculate the crack tip velocity as a
function of applied load, in the regime of steady state notch like crack growth. Our
results are illustrated in Fig. 12, and are in excellent qualitative agreement with the
standard empirical crack growth law v = B e x p (CKI ). Our calculations appear
to underestimate values of C and overestimate B, however. This may partly be
due to inaccuracies in values used for material data, but may also be due to the
limitations of a linear elastic continuum analysis.

6. Acknowledgements
This work was supported by the NIST/ATP membrane program and the MR-
SEC program of the National Science Foundation under award DMR-9632524 with
Brown University.

7. References
Charles, R.J. and Hillig, W.B., (1962), The Kinetics of Glass Failure by Stress
Corrosion, in Symposium on Mechanical Strength of Glass and Ways of
Improving it, Union Scientifique Continentale du Verre, Charleroi, Bel-
gium, pp.511-27.
Chuang, T.-J. and Fuller, E.R., (1992), Extended Charles-Hillig Theory for Stress
Corrosion Cracking of Glass, J. Am. Ceram. Soc., 75 (3) pp.540-45.
348 Z. TANG, A. BOWER AND T.-J. CHUANG

Freiman, S.W., White, G.S. and Fuller, E.R., Jr.(1985), Environmentally Enhanced
Crack Growth in Soda-Lime Glass, J. Am. Ceram. Soc., 68 (3) pp.108-112.
Gehrke, E., Ullner, C. and Hahnert, M., (1990). Effect of Corrosive Media on
Crack Growth of Model Glasses and Commercial Silicate Glasses Int. J.
Glass Sci. Tech. 63 (9) pp.255-65.
Hillig, W.B. and Charles, R.J., (1965), Surfaces, Stress-Dependent Surface Reac-
tions and Strength, in High Strength Materials, V.F. Zackaray. Wiley &
Sons, New York, pp.682-705.
Michalske, T.A., (1983), The Stress Corrosion Limit: Its Measurement and Im-
plications, in Fracture Mechanics of Ceramics, Vol. 5, ed. R.C. Bradt,
A.G. Evans, D.P.H. Hasselman and F.F. Lange, Plenum Press, New York,
pp.277-289.
Peraire, J., Vahdati, M., Morgan, K., and Zienkiewicz, O.C., (1987), Adaptive
Remeshing for Compressible Flow Computations, J. Comp. Phys. 7 2 ,
449-466.
Simmons, C. J. and Freiman, S.W., (1981), Effect of Corrosion Processes on Sub-
critical Crack Growth in Glass, J. Am. Ceram. Soc.,64 (11) pp.683-86.
White, G.S., Greenspan, D. C. and Freiman, S.W.,(1986), Corrosion and Crack
Growth in 33% Na2 O67% SiO2 and 33% Li2 O67% SiO2 Glasses, J. Am.
Ceram. Soc., 69 (1) pp.38-44.
Wiederhorn, S.M., (1975), Crack Growth as an Interpretation of Static Fatigue, J.
Non-Cryst. Solids, 19(1), pp. 169-81.
Wiederhorn, S.M., Evans, A.G., Fuller, E.R. and Johnson, H., (1974), Application
of Fracture Mechanics to Space-Shuttle Windows, J. Am. Ceram. Soc.,
57, pp. 319-323.
Wilkins, B.J.S. and Dutton, R., (1976), Static Fatigue Limit with Particular Ref-
erence to Glass, J. Am. Ceram. Soc., 5 9(3-4), pp.108-12.
Zienkiewicz, O.C. and Zhu, J.Z., (1987), A Simple Error Estimator and Adaptive
Procedure for Practical Engineering Analysis, Int. J. Numer. Meth. Engng,
24, 337-357.
Zhou, S.J. and Curtin, W.A., (1995), Failure of Fiber Composites: A lattice Green
Function Model, Acta Metall. Mater. 43 (8), pp.3093-3104.
ENERGY RELEASE RATE FOR A CRACK WITH EXTRINSIC SURFACE
CHARGE IN A PIEZOELECTRIC COMPACT TENSION SPECIMEN

Anja Haug
Materials Department
University of California
Santa Barbara, California 93106 USA

AND

Robert M. McMeeking
Department of Mechanical and Environmental Engineering
University of California
Santa Barbara, California 93106, USA

Dedicated to James R. Rice on the occasion of his 60th birthday.

Abstract: The fracture behavior of the piezoelectric material PZT-4 in a compact


tension specimen is modelled. The influence of the electrical field and mechanical load
on the energy release rate and the mode mixity ratio is considered. Free charge
accumulation on the crack surface is enforced in the boundary conditions and a finite
element analysis is employed. The results are discussed in comparison with the results
from McMeeking [1] of a crack free of extrinsic charges. It is found that the free charge
on the crack surface diminishes the influence of the electric field on the energy release
rate. Consequently, it may be difficult to deduce the true values of the crack tip stress
intensity factor and the crack tip field intensity factor in an experiment without knowing
the charge condition on the crack surface.

1. Introduction

The influence of electric field and mechanical loading on a piezoelectric compact


tension specimen is investigated. The methology makes use of the J-integral and so this
paper is very appropriate for a collection of articles dedicated to Jim Rice. In addition,
Jims influence on the second author of this paper (R.M.M.) has been extensive from his
days as a graduate student to the present time. The education, mentorship and guidance
that Jim provided has been of enormous benefit and R.M.M. will always be deeply
grateful for this.
McMeeking [1] has already addressed the subject of the paper in the situation
where the surface of the crack is free of extrinsic charge and he has predicted the effect
349
T.-J. Chuang and J.W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 349359.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
350 A. HAUG AND R. M. McMEEKING

of the electric field on the J-integral [2, 3] in that case. However, it is common for free
charge in the atmosphere to be attracted to the intrinsic charge layers on the surfaces of
polarized bodies. It is of interest to investigate the effect on the J-integral of this free
charge accumulated on the crack surface.
The approach of McMeeking [1] is followed to provide a numerical analysis of a
piezoelectric compact tension specimen of a specific configuration and material as used
in some experiments by Park and Sun [4]. This involves finding a stationary value for
the functional given by

(1)
where W, the electrical enthalpy, is

(2)
sij is the strain tensor, E i is the electric field, Di is the electric displacement, A is the
planar area of the specimen and is the displacement relative to the crack plane of one
of the loading points in the direction of its prescribed applied force F (Figure 1). The
interior of the crack is considered to be a sub-region of the domain A, so that the energy
stored in the crack by the electric field contributes to the total electric enthalpy of the
specimen. In this way, the effect of the crack opening on the capacitance and
piezoelectricity of the specimen is accounted for. It follows that the surfaces of the crack
are not components of the perimeter of the specimen but instead the crack surfaces are
treated as interfaces interior to the domain of the calculation. Appropriate continuity
conditions are enforced across these interior interfaces and this will be described below.
The perimeter S surrounding the problem domain is therefore exterior to the rectangle
BCGH as shown in Figure 1 and does not include the crack surfaces. The mechanical
boundary condition applied to this exterior perimeter is that the traction T i is zero. The
electrical boundary conditions consist of zero free charge on the sides CG and HB of the
specimen and a specified value of the potential at electrodes BC and GH glued to the top
and bottom. The final condition is that F is specified.

Figure 1. A compact tension specimen


ENERGY RELEASE IN COMPACT TENSION 351

The constitutive laws connecting the electric displacement, stress, electric field
and strain are
(3)

(4)
where C ijkl is the tensor of elastic stiffness at fixed electric field, e ijk is a tensor of
piezoelectric coefficents and ij is the dielectric permittivity tensor at fixed strain.
The stationary values of are obtained by simultaneous variation of the
displacements u i and of the potential . Stationary values of generated by variations of
u i and unconstrained other than by the boundary conditions are associated with exact
solution of the governing equations of piezoelectricity for the problem including the
boundary conditions. These equations have been summarized by McMeeking [1].
However in this work as in Ref. [1], the finite element method is used and so an
approximate solution is achieved.
The geometry of the problem analyzed is shown in Figure 1. The crack has length
a=6.9 mm, the ligament from the load points to the back face is given by
b=a+c=20.6 mm. The height of the specimen is 2h=19.1 mm and it has thickness
t=5.1 mm. The load points are positioned at a distance d on either side of the crack
where d is 4.5 mm. The loading device is idealized as a pin of zero diameter exactly
fitting into a vanishingly small hole in the specimen and subject to a load F. This is in
contrast to the real component which has a pin of finite diameter [4]. The material is
piezoelectric, poled in the positive x3 -direction and transversely isotropic about the
poling axis. The specimen is assumed to lie in the x1-x 3 plane as shown in Figure 1. The
needed relationships for plane strain of the specimen are:

(5)

(6)

(7)

(8)

(9)
where C i j , ei j and i j are coefficients from equations (3) and (4) stated consistent with
Voigt notation. The potential on the top surface HG is -V and on the bottom surface BC
is V. The electric displacement D1 is zero on the left and right sides BH and CG, a
condition justified by the high dielectric permittivity of the piezoelectric material
compared with that of air. Only the top half of the specimen is analyzed (Figure 1). On
the symmetry line ahead of the crack, OD, the shear traction and the vertical
displacement u3 are both zero. The potential is zero along the entire line AD. The crack
surface is also free of traction and the electric field is continuous across the crack. The
treatment of the opening of the crack, the electrostatic energy in the crack and the total
charge layer on the crack surface are described below.
352 A. HAUG AND R. M. McMEEKING

2. Finite Element Analysis

The finite element equations were solved as a nonlinear system in which the geometric
effect of the crack opening on capacitance was taken into account. As in the treatment by
McMeeking [1], other geometric and material nonlinearities are ignored. In the finite
element code used for the calculations, 4-node plane isoparametric elements with a 2 by
2 rule for integrating the stiffness matrix are utilized. The mesh is shown in Figure 2.

Figure 2. Finite element mesh for the analysis

The condition imposed on the upper crack surface represents continuity of the
electric field and is given by
(10)
where c(x 1) is the crack opening displacement at position x1 on the crack and is equal to
2u 3 and E 3 (x 1 ) is the electric field in the x3 direction in the material adjacent to the crack
surface. This result comes about because it has been assumed that the total surface
charge density on the crack is zero, which requires that in the x3 direction the electric
field in the crack is equal to the electric field in the material, i.e. there is no jump in the
electric field. The condition implies that enough free charge in the atmosphere is
attracted to the surface of the crack to neutralize any intrinsic surface charge induced by
material polarization. Therefore, it is assumed that the atmosphere carrying the free
charges can penetrate the crack and that accumulation of free charge onto the surfaces
takes place sufficiently fast to quickly neutralize the polarization charges. Since 33 i s
zero on the crack surface, equations (6) and (9) can be combined to give at the crack
surface

(11)

Given unit thickness, a nodal charge equivalent to D3 at node i on the crack


surface can be computed by multiplying D3 by L i , which is an effective length of crack
surface associated with node i. The effective length Li is taken to be half the distance
between nodes i-1 and i+1, defined to be the two neighbors of node i on the crack
ENERGY RELEASE IN COMPACT TENSION 353

surface with node i+1 to the right of node i. Equation (10) and (11) are then combined to
provide

(12)

where Q i is the nodal charge for node i, is the displacement of node i+1 in the x 1
direction, is given a similar interpretation for node i-1, i is the potential of node i
and u 3i is the displacement of node i in the x3 direction. It should be noted that to
achieve the expression in equation (12), the approximation

(13)
has been used to estimate the strain s11 at node i.
An iterative approach is used to solve the finite element equations, including the
condition represented by equation (12). At each iteration, the finite element equations
are solved and the residuals at each finite element node are computed. The residuals for
most nodes are simply the un-neutralized nodal free charge and unbalanced nodal forces.
However, for nodes on the crack surface, the electric residual for node i is given by

(14)

where Q i is the current nodal charge for the node i at this iteration and is computed
directly from the finite element equations given that the displacements and potential at
each node has been calculated for this iteration.
The iteration is carried out by a Newton method to drive the residuals for each
-5
node towards zero. Convergence to solutions which change by less than 10 o f t h e
existing nodal potentials and displacements occurs after only 2 iterations. To initiate
iteration the potential on the crack surface is taken to be zero.

3. J-integral

The crack tip energy release rate is equal to the J-integral where J gives the reduction of
potential energy of the specimen per unit area of crack advance [2, 3]. The form suitable
for piezoelectric materials is

(15)

where n i is the normal vector to the contour which completely encloses the crack tip
[1]. As a result, J is the sum of the contributions in the crack where the contour passes
through it and the contribution from the contour in the piezoelectric material. The
contribution to J arising from the segment of the contour within the crack is - W c , since
354 A. HAUG AND R. M. McMEEKING

E 1 and i j are both zero in the crack. This contribution is therefore the negative of the
product of the electrical enthalpy and the crack opening displacement c
where the contour passes through the crack. Let JM be the contribution to J from the
contour through the piezoelectric material. It follows that

(16)

where c is the potential on the surface at the point where the contour integral enters the
crack. JM is computed by the domain integral method as described in [1]. This numerical
evaluation of J is done after the converged finite element solution is obtained. The
following intensity factors are also calculated [1] in the numerical evaluation, defined in
the following way. The Mode I stress intensity factor K I is such that the asymptotic
behavior of the tensile stress ahead of the crack on the crack plane is

(17)

where x 1 is the distance from the crack tip. The electric displacement mode intensity
factor K D is such that the asymptotic behavior of the D 3 -component of the electric
displacement ahead of the crack on the crack plane is

(18)

and the electric field mode intensity factor KE is such that the asymptotic behavior of the
E 3 -component of the electric field ahead of the crack on the crack plane is

(19)

4. Results

The material properties are chosen to be consistent with PZT-4 [1,4]: Elastic Constants
10
(Pa): C 11 = 13.9 x 10 10 , C 12 = 7.78 x 10 , C 13 = 7.43 x 10 10, C 33 = 11.3 x 10 10 ,
C 44 = 2.56 x 1010 , Piezoelectric Coefficients (C/m2 ): e31 = -6.98, e 33 = 13.84, e 15 = 13.44,
-9
Dielectric Permittivities (F/m): 11 = 6.00x 10 , 33 = 5.47 x 10 -9 . The following results
are only valid for the specified geometry (Figure 1) and material data except that values
can be generalized if ratios of parameters are held fixed [1].
In Figure 3 the energy release rate is plotted against the applied electric field Ea
which is computed as V/h. The electric field is made dimensionless by multiplication by
and the energy release rate is normalized by G(F,0), its value at zero applied
electric field. Results are plotted in Figure 3 for 7 values of the applied load F as
ENERGY RELEASE IN COMPACT TENSION 355

indicated in the figure. The results show that an applied electric field reduces the energy
release rate slightly if modest mechanical loads are applied whether the field is positive
or negative. The crack tip energy release rate is independent of the direction of the
electric field. Applying a mechanical load of 5 kN and an electric field in the range
2.3 MV/m in this particular compact tension specimen causes G to differ from G(F,0)
by less than 1.6%.

Figure 3. Crack tip energy release rate for a piezoelectric compact tension specimen in the presence of an
applied electric field and a mechanical load divided by the crack tip energy release rate at the same
mechanical load but without the applied electric field. The results are shown as a function of the applied
electric field and each curve represents a different level of mechanical loading.

Figure 4A and Figure 4B also give the energy release rate, but now plotted
against the applied electrical field divided by the applied mechanical load. (This
parameter is also made dimensionless by multiplication by suitable quantities). The
effect of electric field is almost undetectable for higher forces on the scale used in Figure
4A. However on the range and scale used in Figure 4B, the results indicate that at a
given ratio of electric field to mechanical load, the electric field reduces the energy
release rate by a bigger fraction when the mechanical load is high. It should be noted that
the results indicate that generally an electric field of a few MV/m is required to cause a
change to the energy release rate comparable with 10%; a field of 2.5MV/m and a load
1kN cause a difference of 7.5%. The value of the energy release rate in the absence of
applied electric field is found to be

(18)

where t is the thickness of the specimen. This is identical to the results found by
McMeeking [1].
356 A. HAUG AND R. M. McMEEKING

Figure 4. Crack tip energy release rate for a piezoelectric compact tension specimen in the presence of an
applied electric field and a mechanical load divided by the crack tip energy release rate at the same
mechanical load but without the applied electric field. The results are shown as a function of the applied
electric field divided by the mechanical load and each curve represents a different level of mechanical
loading. A and B show the results in different ranges and to different scales.

In Figure 5 the ratio K E /K I is plotted versus the ratio of applied electric field to
applied mechanical load. Appropriate normalization is used. It can be seen (if one looks
closely at the figure) that there is a non-zero value of K E e q u a l t o a b o u t
even when there is no electric field applied to the specimen. This is
equivalent to a value of K D equal to about 1.1 Thus neither form of the
electrical intensity factors are zero when the applied electric field is absent. However,
the electric displacement intensity factor is almost exactly what is expected from the
nonzero electric displacement induced on the uncracked ligament by the mechanical load
in the absence of applied electric field due to the piezoelectric effect. The nonzero
electric field intensity factors at zero applied electric field therefore arises from this
consideration.
ENERGY RELEASE IN COMPACT TENSION 357

Figure 5. Crack tip mode mixity for a piezoelectric compact tension specimen in the presence of an applied
electric field and a mechanical load. Mode mixity is defined as the crack tip electric field intensity factor
divided by the Mode I crack tip stress intensity factor. Each curve represents a different level of mechanical
loading.

In Figure 5 all curves are indistinguishable which means that a fixed ratio of
applied electric field to applied load always results in the same K E /K1 ratio. The results
for K E and K1 can be summarized by

(19)

(20)

where the slight nonlinearity in the results has been ignored. The result in equation (19)
is identical to that found by McMeeking [1].

5. Discussion

In McMeekings investigation of the energy release rate for a compact tension specimen
[1], he found that when the capacitance of the space in the interior of the crack is
accounted for, the applied electric field has much less influence on the energy release
rate than when the space in the crack is considered to be impermeable to the electric
field. In this paper, we have now found that when the capacitance of the crack interior is
accounted for and it is assumed that free charge in the atmosphere accumulates quickly
on the surface of the crack, attracted there by the intrinsic charge layer induced by
polarization, the effect of the applied electric field on the energy release rate is reduced
358 A. HAUG AND R. M. McMEEKING

even more. This insight can be confirmed by comparing Figures 3 and 4 with equivalent
diagrams in Ref. [1]. However, the effect of the applied electric field is not completely
negligible in this regard and we find that the energy release rate is still influenced
noticeably by the applied electric field.
To a good approximation, the applied electric field has no influence on the
Mode I stress intensity factor and this result is independent of the behavior of free charge
on the crack surface; e.g. KI is the same whether free charge accumulates on the free
surface or not. Moreover, we find that with the conditions imposed in our analysis of the
compact tension specimen, the electric field mode intensity factor is (to a good
approximation) linearly dependent on both the applied load and the applied electric
field. This is in contrast to the results given in Ref. [1] (where free charge does not
accumulate on the crack surface) in which the electric field mode intensity factor was
found to be linearly dependent on the applied electric field but also has strong nonlinear
dependence on the applied load. This behavior observed in the absence of free charge
accumulation on the crack was brought about by the crack opening induced by the
applied load which changes the cracks capacitance. The results of our new calculations
with free charge accumulating on the crack surface indicate that the free charge
diminishes the effect of the changing capacitance as the crack opens due to increasing
applied load. However, the free charge does not make the crack invisible to the electric
field and intensification of the electric field around the crack tip occurs, with the
intensity factor proportional to the applied electric field.
It will have been observed that there is a non-zero electric field mode intensity
factor when the applied electric field is zero and the applied force is non-zero. This
behavior occurs whether free charge accumulates on the crack surface or not and was
also observed in the results of McMeeking [1]. This feature can be attributed to a
piezoelectric effect that is observed to require a non-zero singular electric field and
electric displacement at the crack tip even when there is no applied electric field. In this
sense, the linear dependence of the electric field mode intensity factor on K I observed in
equation (20) is parasitic on the presence of a stress singularity at the crack tip. It is of
interest that a conjugate effect (i.e. a non-zero Mode I stress intensity factor at zero
applied force when the applied electric field is non-zero) is absent or at least negligible
in the numerical results we have obtained.
We concede that our results do not shed any light on why positive electric fields
transverse to the crack in poled PZT4 encourage crack growth and negative ones
discourage crack growth [4-12]. In the results in our paper, the energy release rate is
quadratic in the applied electric field. This implies that if fracture toughness is
independent of the ratio of K E / KI then both negative and positive electric fields should
discourage crack growth. However, as argued in Ref. [1], it is much more likely that the
effective fracture toughness is dependent on the ratio KE /K I providing a possible
explanation of the influence of the sign of the electric field on crack growth. The
effective fracture toughness is the sum of intrinsic and extrinsic contributions, where the
extrinsic contributions include the effect of intergranular residual stresses caused among
other sources by domain reorientation during poling [5, 6] and shielding effects due to
the development of depolarized and repolarized regions of material near a growing crack
tip [1, 9, 10, 12]. Due to the possibility of nonlinear depolarization and polarization
ENERGY RELEASE IN COMPACT TENSION 359

rotation, the extrinsic contribution to the fracture toughness will certainly be influenced
by the direction of the electric field relative to the initial polarization of the piezoelectric
material. Therefore, the dependence of fracture on the sign of the electric field can be
attributed to these phenomena. The results presented in this paper may have a role to
play in the resolution of this issue since it is known that free charge does accumulate on
the polarized surfaces of ferroelectrics [13]. However, the contribution of these results
will be in quantifying properly the stress and field intensity factors for a cracked
specimen in a given experiment. If free charge accumulation occurs quickly on the crack
surface in such an experiment, the field and stress intensity factors must be calculated
along the lines given here so that the toughness can then be identified properly. It should
be noted that the time in which the polarization charge is neutralized is not addressed in
this work. However, we assume that the charge neutralization process is sufficiently fast
that in an experiment it would have occurred prior to the testing of the specimen.

Acknowledgement

This research was supported by Grant 9813022 from the National Science Foundation.

References

[1] R.M. McMeeking, Crack tip energy release rate for a piezoelectric compact tension specimen,
Engineering Fracture Mechanics, 64 (1999) 217-244.
[2] J.R. Rice, A path independent integral and the approximate analysis of strain concentration by notches
and cracks, Journal of Applied Mechanics, 35 (1968) 379-386.
[3] G.P. Cherepanov, Mechanics of Brittle Fracture, McGraw-Hill, New York, 1979, p. 98.
[4] S.B. Park and C.T. Sun, Fracture criteria for piezoelectric ceramics, Journal of the American Ceramic
Society, 78 (1995) 1475-1480.
[5] R.C. Pohanka, R.W. Rice and B.E. Walker, Jr., Effect of internal stress on the strength of BaTiO3 ,
Journal of the American Ceramic Society, 59 (1976) 71-74.
[6] R.W. Rice and R.C. Pohanka, Grain-size dependence of spontaneous cracking in ceramics, Journal of
the American Ceramic Society, 62 (1979) 559-563.
[7] K.D. McHenry and B.G. Koepke, Electric field effects on subcritical crack growth in PZT, Fracture
Mechanics of Ceramics, 5 (1983) 337-352.
[8] A.G. Tobin and Y.E. Pak, Effect of electric fields on fracture behavior of PZT ceramics, in Smart
Materials (V.K. Vardan, ed.) Proc. SPIE 1916 (1993) 76-86.
[9] G.A. Schneider, A. Rostek, B. Zickgraf and F. Aldinger, Proceedings of the 4th International
Conference on Electronic Ceramics and Applications, (1994) 1211-1216.
[10] H.C. Cao and A.G. Evans, Electric-field-induced fatigue crack growth in piezoelectrics, Journal of the
American Ceramic Society, 77 (1994) 1783-1786.
[11] C.S. Lynch, W. Yang, L. Collier, Z. Suo and R.M. McMeeking, Electric field induced cracking in
ferroelectric ceramics, Ferroelectrics, 166 (1995) 11-30.
[12] C.S. Lynch, Fracture of ferroelectric and relaxor electro-ceramics: influence of electric field, Acta
Materialia, 46 (1998) 599-608.
[13] B. Jaffe, W.R. Cook and H. Jaffe, Piezoelectric Ceramics, Academic Press, London and New York
(1971).
This page intentionally left blank.
MICROMECHANICS OF FAILURE IN COMPOSITES
An Analytical Study

Asher A. Rubinstein
Department of Mechanical Engineering
Tulane University
New Orleans, LA 70118

Abstract: An analysis of fracture resistance mechanisms in several composite systems


is presented. A description of the basics of the analytical method developed specifically
for the analysis of fracture development in composites is presented as a unified approach
to different composite systems. The method was applied to several composite systems,
including composites formed from a brittle matrix reinforced by unidirectional fibers,
composites consisting of a brittle matrix reinforced by ductile particles, and a metal
matrix reinforced by ceramic fibers. The reinforcement mechanisms in these composites
are based on the formation of a system of restrictive forces imposed on the crack
surfaces by reinforcing components. The region where these restrictive forces are
activated is represented as a line process zone. A classical fracture mechanics modeling
technique was employed, using the process zone concept and small or large-scale
analysis. The distinctive characteristic of the described method is an explicit
consideration in the analysis of the discrete distribution of the reinforcing components
within the composite. The developed methodology allows one to obtain analytical
solutions to the representative elasticity problems and to investigate detailed
micromechanical aspects of the process.

1. Introduction

Most of the development of analytical methods for the analysis of fracture development
in composites was done in application to ceramic matrix composites (CMC). The
development of these composites has become a topic of significant research effort in
industrial and academic laboratories. The purpose is to take advantage of the
thermomechanical properties of ceramics and to compensate for their low fracture
toughness. The basic ideas explaining the fracture process and effectiveness of fiber
reinforcement were developed using mechanical models of the process. The
development of the methodology for analysis of these composites and relating the
composite microstructure to the fracture resistance was done by Aveston, Cooper and
Kelly (1971), Rose (1987), Budiansky et al.(1988), Budiansky and Amazigo (1989),
361
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 361384.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
362 A. A. RUBINSTEIN

Marshal and Cox (1987), Pagano and Dharani (1990), Rubinstein and Xu (1992),
Rubinstein (1993,1994), Budiansky and Cui (1994), and Meda and Steif (1994a, 1994b)
for fiber reinforced ceramics; Budiansky et al.(1988), Sigl et al.(1988), Erdogan and
Joseph (1988), Bao and Hui (1990), and Rubinstein and Wang (1996, 1998a) for
particulate-reinforced composites, and Rubinstein and Wang (1998b) and Wang (1998)
for metal matrix composites 1 . Although the developed models sometimes reflect
different aspects of the modeled material and employ different analytical and numerical
techniques, they all have a common feature.
The basic physics of the fracture resistance mechanism is the formation by the
reinforcing components of a crack opening constraint in the form of a bridging zone.
The common methodology of composite failure modeling is based on relating the macro
loading parameters to the governing microscale factors, incorporating into the analysis
the effective fiber constraint on crack surface separation. The analytical approach may
differ in consideration of fiber action, using smeared fiber action formulation or
considering a discrete fiber or particle distribution within the bridging zone. The crucial
item for determining the strength and fracture resistance of the composite is the value of
local stress intensity factors acting on internal and surface micro- and macrocracks. The
models have to include the basic material information. The critical information for CMC
fracture process modeling is the effective force imposed by the reinforcing components
on the surfaces of the developing cracks. The fibers or reinforcing particles act as
bridges connecting the crack surfaces and, thus, restricting the crack opening. The
relationship between the force induced by the reinforcing components and the
magnitude of corresponding crack opening displacement plays a key role in developing
the fracture resistance mechanism in the composite. Naturally, this information should
be obtained from experimental observations, and it should correspond to a specific
composite system. A number of factors influence this relationship, and a number of
procedures for obtaining the data could serve as the basis for determination of the fiber
pullout - force relationship, F(u), for a specific composite system, or ductile particle
deformation pattern within the bridging zone. Several of these relationships and their
effect on fracture resistance development were investigated.
The analytical approach described in this paper is based on a discrete distribution of
the reinforcing components within the bridging zone; this is a distinctive quality of the
method. In this approach, a model is developed based on an exact analytical solution of
the corresponding problem that represents the processes taking place during the crack
growth. Using this approach, all fracture mechanics parameters can be monitored at any
intermediate step as the crack tip progresses between the reinforcing particles or fibers.
The main advantage of this methodology, as compared to the methods based on
smearing of the fiber or particle action within the bridging zone, is obtaining actual
values of all fracture mechanics parameters rather than their average values. The
numerical procedure required for parametric study of the model also appears to be
simpler and more direct. In both cases of discrete fiber or particle distribution and
smeared reinforcing action within the bridging zone, the process zone is treated as a line
zone ahead of the crack, and the elastic properties of the composite outside of the
1
The reference list on the subject is not complete; only principal references are cited.
MICROMECHANICS OF FAILURE IN COMPOSITES 363

process zone are assumed to be homogeneous, although not necessarily isotropic.


In the following sections, the basics of the analytical formulation as applied to cases
of reinforcement by unidirectional fibers, ductile particle reinforcement of ceramic
matrix, and reinforcement in metal matrix composites will be outlined, and several
examples of the results will be described. The emphasis here is on the similarities in
mathematical formulation as applied to different physical systems.

2. Basics of the Modeling Scheme

The process of fracture resistance development requires a three dimensional description.


Considering the periodic distribution, the analysis is formulated for a two dimensional
plane which is placed through the centers of the particles, or fibers, perpendicular to the
crack plane and aligned with the loading direction. The developing crack front takes the
wavy form, as observed by Botsis and Shafiq (1992). The maximal values of the stress
intensity factors along the crack front will be on the trailing portions of the crack front.
The analytical formulation described here corresponds to the family of planes
intersecting locations of maximal stress intensity along the crack front. The stress
intensity factors controlling the crack growth and the most significant effects associated
with the reinforcement are taking place in this plane. Therefore, focusing attention on
this plane is justifiable. The stress state, however, is not exactly in the category of plane
stress or plane strain types. Because the problem is periodic, we assume the stress state
to be close to the plane strain case, and therefore consider it as such. This assumption is
commonly used in the modeling of the fracture process in composites.
The analytical formulation of the model is based on the classical description of the
stress state in an elastic body in terms of analytic potentials (Muskhelishvili (1963)).
The conventional definition of the analytic stress potentials, (z) and (z ), is given by
relationships (1); using standard notations, is a shear modulus, = 3-4v for plane
strain or = (3-4v)/(1+ v) for generalized plane stress, and v is Poissons ratio. The
complex displacement is given in (1) in the form of a gain between two points A and B
on the complex plane, and the complex force in (1) is given as a resultant of traction on
an arc connecting these points.

(1)

Position the crack and the bridging zone along the line y=0 on the complex plane and
consider Mode I type loading. Under these conditions and Mode I loading the symmetry
of the problem allows one to relate the unknown stress potentials as
364 A. A. RUBINSTEIN

(2)

for a small scale analysis. The small scale approach limits the size of the bridging zone
as compared to the size of the crack and any other geometrical parameter of the
problem, thus assuming the crack to be infinitely large. In case of finite crack size and a
comparable bridging zone, relationship (2) changes as given in equation (3) (Rubinstein
(1994)) with being the applied stress.

(3)

Thus, the problem is reduced to one unknown function. Consideration of Mode I - type
loading does not limit the generality of the analysis. The corresponding results for Mode
II and Mode III may be obtained from the results derived for Mode I. Both of these
cases also are reduced to one unknown function with the same boundary conditions as in
Mode I. For example, in case of Mode II, the relationship between the complex stress
potentials becomes (Rubinstein (1985)),

(4)

Generally, the stress potential '(z) for Mode II type loading can be obtained using the
solution for Mode I by substituting -iK II( ) for K I( ) in potential '(z), where i is the
imaginary unit. Solutions for different fracture modes are similar up to the conditions on
the fibers or other reinforcing components under consideration. These conditions
usually change under different modes of fracture. For example, if we consider fiber
pullout versus force on the fiber relationship, it is unreasonable to expect this
relationship to remain unchanged under these two different conditions. The system of
forces acting on the fiber and conditions on the fiber-matrix interface change. Therefore,
in case of mixed mode loading conditions on the reinforcing component, the resulting
effect depends on the specific relationship between the loading modes. Usually, Mode I
is considered as a dominant fracture mode in reinforced composites. The components
manufactured from these composites are usually designed to have the reinforcement
aligned in the direction of the maximal load.
The approach to the analysis of brittle matrix reinforcement presented here is based
on a micromechanical consideration. The action of each reinforcing component is
considered as part of the system of activated reinforcing elements within the bridging
zone, and yet they are considered as discretely spaced elements. The distribution of
these elements within the bridging zone may take an arbitrary form, and the analysis
presented here may accommodate it. However, in most cases this distribution is
considered to be periodical; these cases will be presented below. This is also typical for
other models which are based on substitution of the actions of the discretely distributed
reinforcing elements by the set of continuously distributed forces within the bridging
zone. Consideration of the periodically distributed reinforcing elements, using the
methodology presented here, is strictly a matter of convenience; this methodology can
MICROMECHANICS OF FAILURE IN COMPOSITES 365

be applied to any systematic or random distribution.


To describe the boundary conditions, consider here small scale models only. The
large scale cases are developed in a similar manner by Rubinstein (1994). For Mode I
small scale conditions, the complex potential behavior at infinity has to be specified as

(5)

Additional conditions for the analytic function '(z) have to be stated along the x-axis.
Position the crack and the bridging zone along the x axis, y=0, with the bridging zone
from x=0 to x=c, and the crack on - < x < 0. The remote Mode I loading is assumed to
be applied along the y axis. The reinforcing components with thickness a are distributed
with period p, and the first component within the bridging zone is on the interval 0 < x <
a. The thickness of the component could be the fiber diameter or the particle diameter.
Formulation of specific boundary conditions taking place on these intervals for different
types of composites are outlined in the following subsections. Using equations (1) and
(2), the shear stress on y=0 is vanished and the normal stress and the displacement are

(6)

2.1 UNIDIRECTIONAL FIBER REINFORCEMENT

The boundary conditions along y=0 state the stress free crack surface, x < 0, and the
crack ligaments between the fibers. On the ligaments representing the fibers we state
the condition of a constant displacement. This condition was determined from
consideration of the stress state around a cylindrical fiber under the force pulling it out
from the matrix. Under this stress state, the matrix deformation around the fiber will
assume a cylindrical symmetry, and, therefore, the displacement along the rim of the
fiber-matrix interface on the surface will be constant. Importantly, the stress state in the
matrix in the vicinity of the fiber in the plane of consideration corresponds to a nearly
undeformed fiber-matrix interface. These facts led us to the statement of the constant
displacement on the ligaments representing fibers. The constant displacement statement
also enforces the fact that each fiber is pulled out from the matrix on its specific constant
amount along the crack surface intersecting with the fiber-matrix interface. Any
solution to a boundary value problem with more complicated boundary conditions still
will have a solution given below as a homogeneous part of any more general solution.
Thus, using equations (6), the boundary conditions for a bridging zone with N fibers are

f
(7)
366 A. A. RUBINSTEIN

Figure 1. Bridging zone formed by unidirectional fibers.

The corresponding geometry is illustrated in Figure 1. In addition to these conditions a


physical condition relating the fiber pullout displacement, u, and the force on the fiber,
F, has to be introduced into consideration. The relationship F(u) has to be obtained
from an experiment. Marshal and Price (1991), and Carter, et al (1991), reported a
practically linear relationship between the fiber pullout displacement and the force on
the fiber, at least during the initial stage. On the other hand Mumm and Faber (1995),
reported a parabolic relationship. Analytical models used both relationships. Of course,
these data correspond to different composite systems. However, comparing results of the
analysis for equivalent situations, the difference in the pattern of F( u) does not seem to
be as important as the corresponding numerical values of this function. Results shown
here correspond to a linear function, F( u), with a parameter representing interface
properties.

(8)

Subscript k (k = 1, 2,, N ) indicates the particular fiber to which relationship (8) is


applied. This equation completes the set of conditions for the problem. The stress
potential was found in the form

(9)

The N constants d k appearing in equation (9) were found numerically by enforcing


condition (8) and solving a system of linear equations. Function (9) was formed on the
basis of the result of the Keldysh - Sedove problem (Muskhelishvili (1963)).
Computation of the force and displacement components for equation (8) was conducted
MICROMECHANICS OF FAILURE IN COMPOSITES 367

using the Gauss - Chebyshev quadrature. Solution (9) is a solution to a homogeneous


problem; it is a dominating part of any other solution with different boundary conditions
on the ligaments representing the fibers.

2.2 PARTICULATE REINFORCEMENT OF A BRITTLE MATRIX.

The toughening mechanism in these composites is based on an attraction of the crack tip
to particles of a lesser stiffness. Some aspects of the crack attraction to defects were
investigated by Rubinstein (1986). The ductile particles are forming plastic bridges as
the crack front passes through. The extensibility of these bridges depends on the
particle ductility and the strength of the particle-matrix interface. The interface strength
controls the development of the particles shape during the plastic deformation and thus
influences the tri-axial stress state within the particles. A detailed numerical study of
developing particle shapes was reported by Tvergard (1992, 1995), the experimental
observations were reported by Venkateswara Rao et al. (1992), and plastic deformation
development in constrained long wires was reported by Ashby et al. (1989). The
theoretical analysis presented here departs from the traditional methodology developed
for these materials by Budiansky et al. (1988), Erdogan and Joseph (1989), and Bao and
Hui (1990). The presented analysis, Rubinstein and Wang (1996,1998), departs from the
traditional approach of smearing the action of the particles over the bridging zone using
continually distributed forces. The particles here are considered to be distributed
periodically, with period p, the active cross section of the particle k within the line
bridging zone is between points ak and b k , when k is counted from the beginning of the
bridging zone at x=0; this geometry is illustrated in Figure 2. The normal stress along
the crack line in the plastically deforming particle k is k , when the deformation within
the particle is fully plastic, and FY when the plastic zone initiates. The particles are
assumed to be ideally plastic with yield stress Y . The particles from a strain hardening
material may be included in this analysis without additional effort by adjusting values k
according to the extent of deformation of each particle. In the presented work, the intent
of considering the different values of the plastic stress on each particle was to include
the tri-axiality effect. Thus, the boundary conditions for the stress potential, '( z ), for the
corresponding plane problem are

(10)
368 A. A. RUBINSTEIN

Figure 2. Bridging zone formed by ductile particles.

The shape evolution of initially spherical particles during the intense plastic deformation
is approximated as a neck of parabolic profile, (11),

(11)

which connects the intact portions of the particle-matrix interfaces, yk here is the vertical
coordinate of the intersection of the neck with the intact portion of the spherical particle-
matrix interface, and x k is the corresponding horizontal coordinate. Parameter A is
introduced as a characteristic of the interface strength. A weak interface corresponds to
A=0, and a strong interface corresponds to high values of A. Several examples of the
particle profiles with different interface strength are given in Figure 3. The radius of the
narrow cross section, in this case is given by equation (12).

(12)

So a k = pk-r k , and b k =pk+r k . Additionally the condition of a constant particle volume


during the plastic deformation is enforced. Using average crack opening displacement,
u k , on k-th particle, this condition on particle k is given by equation (13)

(13)
MICROMECHANICS OF FAILURE IN COMPOSITES 369

Figure 3. Particle deformation shapes for different interface strength parameter A. A=0 corresponds to weak
interface, A=1, .., 10 - intermediate, and A=50 - strong interface.

The solution of the boundary value problem (10) is given by equation (14).

(14)

Integrating function (14), one finds the corresponding potential (15) and displacement
using equation (6).

(15)
370 A. A. RUBINSTEIN

The values of rk , a k , b k and u k were found numerically by solving the nonlinear system of
equations (11), (12) and (13) simultaneously, thus basically solving a system of
nonlinear algebraic equations, rather than dealing with nonlinear integral equations as
traditional methods would require. A special case, when A =0, corresponds to a
cylindrical shape of deforming particle; the computations are significantly simplified in
this case because equations (12,13) can be solved analytically. Generally, the numerical
procedure involves evaluation of the current crack tip position and adapts according to
its position. Thus, depending on whether the crack tip is located between the particles,
or within the particle, or depending on the extent of plastic zone development, a
different algorithm is applied, Rubinstein and Wang (1998). Those algorithms are based
on specific conditions for the crack growth appropriate for the considered interval. If the
crack tip is located between the particles, the constant stress intensity factor equal to the
critical value for the matrix, K mIC has to be maintained; if the crack tip is located within
the particle, different conditions, depending on the particle size, must be applied. To
relate the particle size and other mechanical parameters of the composite system a
parameter k was introduced (16).

(16)

In fact this parameter relates the particle size to the Dugdale plastic zone size in the
material with the same yield stress under an acting stress intensity factor equal to the
matrix toughness, K mIC . When k=1, the diameter of the undeformed particle is equal to
this Dugdale zone size, and in this case, as well as in case k<1 when the crack tip
reaches the particle, the plastic yield within the particle will develop throughout the
particle without necessarily increasing the remote loading. In other cases the plastic
zone will initially develop, but to continue throughout the particle, the remote load has
to be increased. In all cases it was found that the point of highest resistance to future
crack advancement is the instant when the crack tip advances from the particle into the
matrix, Rubinstein and Wang (1998).

2.3 METAL MATRIX REINFORCED BY CERAMIC FIBERS

Reinforcement of a metal matrix by means of aligned fibers or ceramic layers can be


very effective in improving the fracture resistance of base metals, Bloer, et al. (1996).
The model of the fracture resistance mechanism development in these composites is
developed here. The aim is to determine the relationships between the composite
parameters and fracture resistance effectiveness. This method outlined above was
advanced to allow nonlinear behavior of the matrix within the bridging zone, making it
applicable to a broad class of composite materials. As in previous cases, a discrete
distribution of the reinforcing components is considered, and the model is based on an
exact analytical solution of the corresponding boundary value problem.
Within the line process zone, the following effects are taking place. The intact fibers
MICROMECHANICS OF FAILURE IN COMPOSITES 371

(or elastic layers) bridge the failed, partially failed, or plastically deformed matrix. We
consider the area between the fibers where the matrix may experience plastic
deformation. The intensity of plastic deformation may be significant to initiate void
formation on that interval. The interval is classified as partially failed if the void size is
less than the space between the fibers. If the void reaches the size of the space between
the fibers, we consider this region as failed. The plastic flow of the matrix is constrained
by the elastic fibers. Therefore, the spread of the plastic deformation from the crack
plane cannot be anticipated to be significant. This justifies our assumption of a line-like
plastic zone, in a way similar to the Dugdale model for a single crack (Dugdale (1960)).
As a criterion of the matrix failure, or as a void initiation criterion, we use a critical
value for the crack opening displacement, c.

2.3.1 Physical Aspects and Assumptions of the Model.


The physics of the process zone formation and development requires a special
consideration. A few aspects of the reinforcement mechanism are not typical for other
fiber-reinforced composites. One of these aspects deals with special properties of the
relationship between fiber pullout displacement and the force on the fiber, and another
aspect deals with the process of void formations between the fibers. In both cases the
simplest physically admissible conditions are applied.
The fiber pullout relationship is assumed to be linear, in the following form:

(Fi F 0i )= B i i = 0,1,, N (17)

Here is again the pullout coefficient, Fi and B i are force and displacement
respectively on the (i +1)th fiber, and F0i is the threshold force on this fiber. One must
introduce the threshold force, F0i , on the fiber, to assure consistency of the physical
processes in this composite system. The threshold force will simply prevent matrix
separation before plastic deformation in the surrounding fiber area takes place. We
introduce F0i as a force on the (i+1)th fiber produced by the surrounding stress field,
which initiates the plastic flow around that fiber. In practical terms, we use the value of
the force corresponding to the stress state when the plastic zone tip just passes the fiber.
This value of the threshold force is physically suitable, although it is not the only
possible definition. There is no specific experimental information on what value should
be used for the threshold force.
Another aspect of the model requiring special treatment is the void formation
process. The initiation of the void formation is assumed to be the state when the crack
opening displacement within the plastically deformed area reaches the critical value.
The expected location is on the interval between the fibers, somewhere near the center
of the interval. The yield stress acting within this region acts as surface traction on the
crack surfaces. As soon as the void is initiated, the traction is removed from the voids
newly formed free surface. The removal of the surface traction from the crack surfaces
increases the crack opening displacement and, in turn, stimulates void growth. The void
size is determined by the critical value of the crack opening displacement, which is
maintained at the ends of the void. This is a typical instability of the physical process,
372 A. A. RUBINSTEIN

which must be reflected in the model and in the numerical implementation of the
process. Thus, there can be no continuous quasi-static void growth from infinitesimal
size to any intermediate size. The only possible static states are no void or a void of a
finite equilibrium size. The model reflects this aspect of the physical process, and the
numerical scheme is specifically designed to accommodate it. The numerical scheme is
described in detail by Wang (1998). The same criterion using the critical crack opening
displacement at the tip is applied for determination of the main crack position.
The resulting nonlinear model represents all basic physical aspects of nonlinear
behavior of the material within the process zone, including the instabilities associated
with void formation and the development of traction-free crack surfaces. The model
demonstrates the effectiveness of the plastic yield on the fracture properties of
composites.

Figure 4. The process zone geometry and corresponding boundary conditions.

2.3.2 Analytical Development


The analytical formulation of the model is conducted in terms of a two-dimensional
problem as in previous sections. The region surrounding the process zone is treated as
elastic material, and the analysis is conducted using the method outlined above. The
plane of consideration is the plane perpendicular to the crack plane and crack front, and
passing through the centers of periodically spaced fibers. In the case of a layered
composite, it is a plane perpendicular to the layers. The symmetry conditions are used
again, reducing the problem to one unknown analytic potential '(z). The boundary
conditions for this potential are set along the crack line, z = x + i0, as follows:
(a) on the crack and void surfaces Re '(x) = 0, traction free surface condition;
(b) on the segments where the plastic yielding is active Re '(x) = 0 /2, where 0 i s
the yield stress;
(c) on the fibers and in front of the process zone Im'(x) = 0, stating constant or
zero displacement;
MICROMECHANICS OF FAILURE IN COMPOSITES 373

(d) as z : '(z) K /2(2 z)1/2 , the remote loading condition.


The above boundary conditions and geometry of the process zone are illustrated in
Figure 4.
For convenience of development of the numerical procedure, the case of
periodically spaced fibers, or layers, is considered with period p, and uniform fiber
thickness a. This is not an essential assumption; the solution can be easily extended to a
more general case. The function '(z) represents a combination of both cases considered
in previous sections. The function was constructed using the Keldysh-Sedov theorem,
Muskhelishvili (1972). According to this theorem, the boundary conditions outlined
above determine the function uniquely up to the type of singular behavior at the ends of
the intervals determining the real and imaginary parts of the function. Setting this
behavior in a form suitable for the physical conditions of the problem, we construct the
complex potential as

(18)

The function R(z) in (2) is defined as

(19)

The process zone starts at the crack tip position c t ; it includes N fibers and the plastic
zone in front of the N th fiber ending at point d t . The integrals in (18) are taken over all
intervals representing the plastically deformed area and the segments representing voids.
It does not include segments representing fibers. The switching function Tj (t) in (18) is
introduced for computational convenience as Tj (t) = 1 on segments with active plastic
yield, T j (t) = 0 on segments representing voids formed within the plastic region. The
limits of the integral are specified in general terms here to include all the appropriate
intervals. The process zone with the boundary conditions is illustrated in Figure 4. The
function '(z), as given by equation (18), contains N+2 constants Cm , and unknown
limits of the void boundaries on each interval. The first part of (18) represents the
homogeneous solution of the corresponding boundary value problem. To find the
constants and the voids boundaries, one uses the set of equations (17) which apply to
each segment representing the fibers. The net force on the fiber is computed as
374 A. A. RUBINSTEIN

(20)

The crack opening displacement is calculated using the following integral:

(21)

B i in set of equations (17) is the total crack opening displacement calculated for a fiber i
which corresponds to x = a + pi in equation (21).
An additional set of conditions is set to determine the position of the crack tip, c t ,
and the sizes and positions of the voids on each interval between the fibers, g i and h i ,
where i corresponds to a specific interval within the process zone. These conditions are
for all i:

(22)

The displacements in (22) have to be computed using equation (21). If the crack opening
displacement within the space between the fibers on a particular interval does not reach
the critical value, no void will form. The final condition to be satisfied is the condition
of bounded stress at the initiation of the plastic zone in front of the process zone. This
condition is:

(23)

Equations (17), (22) and (23) with formulae (18-21) form a system of nonlinear
equations which were solved simultaneously to obtain constants Cm, positions ct , d t , and
g i with h i. Details of the development of the numerical scheme are described by Wang
(1998).
MICROMECHANICS OF FAILURE IN COMPOSITES 375

3. Fracture Resistance of Composites.

The numerical procedure developed on the basis of the described methodology can
provide a broad range of information about the processes taking place within the
bridging zone of the composites and a complete stress field at any point outside the
bridging zone. Development of fracture resistance in composites and its dependence on
internal microstructure were prime targets of the analysis. A few numerical examples of
the developed models are given below. The emphasis of the selected examples is on the
effects associated with the discrete distribution of the reinforcing components.

Figure 5. Fracture resistance curves for fiber reinforced ceramics. The length of the bridging zone is given in
terms of the number of participating periodic cells.

3.1 CERAMIC MATRIX COMPOSITES.

In Figure 5, the development of fracture resistance curves in ceramics reinforced by


unidirectional fibers is presented. The data are given as a function of a number of
periods with active fibers participating in the bridging zone and as a variation of the
dimensionless fiber pullout parameter . Additionally, the cases of different fiber
spacing are shown in Figure 5. This dependence on the fiber spacing aspect ratio is an
important composite fabrication parameter which controls a variety of properties. The
fracture resistance is measured as an ability of the composite to prevent matrix cracking.
In these terms, a composite with the fibers positioned extremely close to each other will
have no or very little fracture resistance. This result is demonstrated in Figure 6, where
376 A. A. RUBINSTEIN

fracture resistance is given as a function of the fiber spacing aspect ratio for a given
number of fibers participating in the bridging zone. A similar pattern was observed in
the analysis of the composite strength, Rubinstein ((1994). In the cases discussed here
of fiber reinforced brittle matrix, a linear relationship (8) was used describing the
relationship of fiber pullout displacement to the force on the fiber. However, this is not a
limitation of the described method. Several nonlinear cases were investigated for the
simplified case of one bridging fiber, Rubinstein (1993). The advantage of considering
that case was availability of a complete closed form analytical solution. The results
demonstrated that this nonlinear pattern of this relationship plays an important role only
when the crack tip is located very close to the fiber. Thus, it would apply to cases of
very high volume fraction of fibers. Otherwise the linear pattern gives very similar
results.

Figure 6. Resistance curves vs. fiber spacing aspect ratio a/p; the length of the bridging zone is constant.

3.2 PARTICULATE-REINFORCED CERAMICS

The results illustrating the behavior of the particulate-reinforced ceramic matrix are
based on the analytical scheme which considers discrete particle distribution. However,
to relate the obtained results with other methods based on smeared continuous force
distribution within the bridging zone, the average stress versus average displacement are
given in Figure 7. The data in Figure 7 were obtained using a discrete particle
distribution. The fracture resistance of particulate-reinforced composites depends on a
number of parameters: particle spacing, particle size and ductility in relationship to
matrix toughness, and the strength of the particle-matrix interface, to name a few. The
developed analytical and computational schemes allow one to investigate every aspect
MICROMECHANICS OF FAILURE IN COMPOSITES 377

of the processes taking place during the crack growth and formation of the bridging
zone. The particle size and yield stress form a parameter (16), k, which if related to the
matrix toughness determine whether the plastic zone will develop all the way through
the particle immediately after the crack tip has approached it, or the particle may
partially yield and require a significant additional load in order for the crack front to
pass through. The particle spacing is important for fracture resistance development. The
closer the particles are positioned, the more energy is absorbed due to plastic
deformation.

Figure 7. Average bridging stress as a function of crack opening displacement and the interface strength
parameter

Figure 8. Fracture resistance curves for particulate reinforced ceramics.


378 A. A. RUBINSTEIN

The particle-matrix interface plays a special role in this process. The quality of the
interface determines the pattern of the plastically deforming particles. The maximal
length of the deformed particles controls the potential length of the bridging zone, and
the changes in diameter of the particles determine the load carrying capacity of the
particles. Thus, it is not surprising that the composites with a weak interface exhibit
higher fracture resistance, Rubinstein and Wang (1998). These results are consistent
with experimental observations reported by Venteswara Rao et al (1992). However, the
very weak interface will allow the crack to pass through by debonding the particle-
matrix interface before the particles exhibit substantial plastic deformation. These cases
were not included in the presented analysis. The limit of the effectiveness of the
interface has to be evaluated. The fracture resistance data corresponding to a weak
interface are presented in Figure 8. The data in Figure 8 include dependence of the
fracture resistance on particles spacing aspect ratio d/p; these data are converted to
particles volume fraction as well.

Figure 9. Fracture resistance curves.

3.3 METAL MATRIX COMPOSITES.

To demonstrate the effectiveness of metal matrix reinforcement, the fracture resistance


curves were computed. The fracture resistance curves shown in Figure 9 are normalized
by the stress intensity factor K0 . This is a stress intensity factor which would generate a
Dugdale zone of a maximal size for a specified critical value of the crack tip opening
displacement, c , in the matrix without reinforcement. The critical value of the crack
opening displacement, c, and dimensionless fiber pullout parameter, , are important
MICROMECHANICS OF FAILURE IN COMPOSITES 379

parameters controlling the degree of fracture resistance development. is defined in


equation (24).

(24)

Because the numerical values of these parameters are not available for the material
systems described in the literature by Bloyer, et al. (1996), we may compare our results
with the experimental data qualitatively only. The data depicted in Figure 9 represent
three cases, corresponding to two sets of value of , and two sets of critical values of
crack opening displacements. The R-curves in Figure 9 are given as functions of the
position of the leading end of the process zone normalized by the period of the
composite system structure. This form of presentation is important because it provides
information on the extent of process zone development or damage surrounding the crack
tip. However, the experimental measurements would most likely be conducted as a
function of the crack tip position. Although qualitatively the results will appear to be
similar, quantitatively they may differ. The developed model can supply information for
both cases. In Figure 9, the line breaks correspond to the instances when the front end of
the line plastic zone passes reinforcing fibers. Only relatively short bridging zones were
investigated.
Reinforcement of the metal matrix appears to be very effective in improving
fracture resistance and in constraining the plastic flow typically surrounding the crack
tip in ductile material.
The values of parameter were chosen without a relationship to a particular
composite system. As discussed in previous sections, the fiber pullout displacement has
to be compatible with the displacement due to the plastic flow and constraint by the
surrounding elastic material. Contrary to fiber reinforced ceramics, not all values of
are admissible to a particular composite system. The limitations of these values and the
relationship of these limitations to other composite parameters are not completely
understood at this time.
The modeling technique employed in this investigation is new and makes available
for evaluation certain features of the physical process that are not available if other
methods of analysis are employed. For example, if the method of smearing of the active
forces within the bridging zone is applied, and on that basis the method of integral
equations is employed, the basic relationship of the average bridging stress within the
process zone versus crack opening displacement must be specified. These relationships
may be derived from a simplified cell model. The method developed for this study
allows us to obtain the actual relationship within the investigated process zone, simply
by following one period cell after it has entered the process zone. The resulting
relationship is depicted in Figure 10.
This relationship, as was expected, is highly nonlinear, and contains softening
regions due to void formation within the process zone. The softening regions are shown
as straight dashed lines; however, this is not the actual pattern of the process. As
discussed earlier, at the moment a void is initiated, the system loses stability. The next
stable point is the instant when the void reaches equilibrium size. The space between
380 A. A. RUBINSTEIN

these two states was connected on the graphs with straight lines to maintain continuity
of each considered case. Accommodation of these unstable intervals during the process
zone development represented a challenge for development of the numerical procedure
for this analysis.

Figure 10. The average stress versus average COD relationship within the process zone.

In Figure 11, a typical crack opening displacement pattern within the bridging zone and
its vicinity is shown. The fibers are indicated on Figure 11; the black regions represent
the plastically deforming regions. The lighter areas on the fibers indicate the portion of
the fiber pulled out from the matrix according to equation (17). The voids depicted in
Figure 11 were developed according to the outlined process. The computational
procedure required evaluation of the crack opening displacement at every intermediate
step. Observation of the development of the process zone was very helpful in
understanding the physical process, and as a controlling factor for the numerical
procedure. As was observed, the voids due to the described unstable growth process
have a tendency to spread over the space between the fibers rather quickly, leaving only
narrow regions adjacent to the fibers where plastic yield takes place. The example
presented in Figure 11 corresponds to the case of a/p = 0.1, = 0.2, and c /p = 0.2.

4. Concluding Remarks

A unified approach to the failure analysis of different composite systems was presented.
The methodology outlined here offers a powerful tool for micromechanical analysis of
composite materials. Using this approach, one may obtain very detailed information
MICROMECHANICS OF FAILURE IN COMPOSITES 381

about the processes taking place within the process zone of the composite. Variation of
any micromechanical parameter may be accurately evaluated at any intermediate step of
the process zone development. One of the specific advantages of this methodology is
explicit participation in the analysis of the spacing of the reinforcing components. The
interaction effects of these components participated in the stress analysis in their
vicinity. These details may be used for determination of potential debonding, in the case
of fiber composites, or crack path deflection along the fiber-matrix interface. In the case
of particulate reinforcement, the quality of the particle-matrix interface plays a principal
role in the developed analysis, which allows one directly to examine the associated
effects. All details of the ongoing processes were accounted for within the process zone
of the metal matrix composites. The considered processes included elastic deformation
of the fibers, the relationship between fiber pullout and force on the fiber, plastic
deformation of the matrix, formation of voids and associated instabilities. All these
processes are taking place on each period of the fiber spacing within the process zone.
The method does not have a limit for the number of periodically spaced fibers
considered.

Figure 11. Crack opening displacement within the process zone and its vicinity resulting from numerical
simulation. Black regions indicate plastically deformed regions.

In all considered cases, exact solutions to the corresponding boundary value


problems were formed, and the constant parameters involved in the solutions were
obtained using computational schemes. The computations were organized in an error-
controlled environment wherein the maximal acceptable cumulative error was specified
in advance.
The numerical procedures formulated for the described approach are simple and
382 A. A. RUBINSTEIN

based on the solution of a system of algebraic equations with the number of unknowns
equal to the number of reinforcing components involved in the developed process zone.
This is contrary to traditional schemes based on solution of integral equations where the
numerical discretization of the interval usually requires a substantial number of nodes in
order to obtain reliable accuracy of the results.
The presented analysis was developed using methods of two dimensional elasticity.
The results obtained here are projected to a three dimensional problem characterizing
composite behavior. This is a legitimate projection for the following reasons. The
analysis of the bridging mechanism consists of two parts. These are the stress analysis of
the region surrounding the bridging zone, and the analysis of the constraint imposed by
the composite reinforcing components. The two parts must be matching, consistent, and
compatible. The stress analysis developed here represents an exact solution of the
corresponding problem applicable to a specific plane passing through the centers of the
fibers or inclusions, and which is perpendicular to the crack front. This plane is a plane
of symmetry for the considered periodic problem. The average strain over the period in
the direction perpendicular to the considered plane is zero. Therefore, the plane strain
equations can be applied to obtain an average result. The crack front, on the other hand,
is not straight. It develops a wavy pattern as the crack front progresses from one
reinforcing array to the next. The stress intensity factor acting within the considered
plane plays a decisive role in determining the future crack progress. The plane problem
implies a straight crack front and, thus, the results of the analysis may be used as a
conservative prediction of fracture resistance development. The stress functions (9), (14)
and (18) represent the stress state on the described plane, and they, of course, could be
directly applied to the fracture analysis of layered composites. On the other hand, the
constraint on the crack opening displacement imposed by the reinforcing components
was constructed specifically for the considered composite materials. The displacement
restriction on the fibers for fiber reinforced composites, second equation in system (7),
the particle shape development for particulate reinforced composites, equations (12) and
(13), and equation (17) for metal matrix composites, are based on local cylindrical
symmetry. In the case of particulate composites, maintaining constant volume of the
plastically deformed material in each spherical particle restricts the crack opening
displacement. Thus, the crack opening displacement restrictions brought in through the
conditions on the fibers or the plastic particles are the key components relating the stress
functions to the specific cases considered here. To apply the developed analysis to
layered composites, one would have to restructure the displacement conditions imposed
by the reinforcing components.
The presented analytical method of composite analysis relates essential composite
design parameters with composite performance in terms of fracture resistance, thus
relating essential micromechanical processes taking place on microscale with composite
performance. This approach can be very useful in optimal composite design.

Acknowledgment This work was supported by National Aeronautic and Space


Administration, Air Force Office of Scientific Research, and in part by Kyoto University
Foundation through a visiting scholar fellowship.
MICROMECHANICS OF FAILURE IN COMPOSITES 383

References

Ashby, M. E., Blunt, R. J. and Banister, M. (1989) FIow characteristics of highly constrained metal wires.
Acta. Metal. 37 (7), 1847-1857.
Aveston, J. Cooper, G. A. and Kelly, (1971) A. Single and multiple fracture, Conference on The Properties of
Fiber Composites, National Physical Laboratory, Guildford, Suerey, ICP Science and Technology
Press, pp. 15-26.
Bao, G., and Hui, C.-Y. (1990) Effects of interface debonding on the toughness of ductile-particle reinforced
ceramics. Int. J. Solids Structures. 26 (5/6), 631-642.
Bloyer, D. R., Venkateswara Rao, K. T. and Ritchie, R. O., (1996) Resistance-Curve Toughening in
Ductile/Brittle Layered Structures: Behavior in Nb/Nb3 Al Laminates. Materials Science and Engineering
A, 216, 80-90.
Botsis, J. and Shafiq, A. B., (1992) Crack growth characteristics of an epoxy reinforced with long aligned
glass fibers, Int. J. Fracture, 58, R3-R11.
Budiansky, B. Amazigo, J. C., and Evans, A. G. (1988) Small scale crack bridging and the fracture toughness
of particulate-reinforced ceramics. J. Mech. Phys. Solids. 36 (2), 167-167.
Budiansky, B. and Amazigo, J. C. (1989) Toughening by aligned, frictionally constrained fibers. J. mech.
Phys. Solids 37, 93-109.
Budiansky, B, and Cui, Y. L. (1994) On the tensile strength of a fiber-reinforced ceramic composite
containing a crack-like flaw, J. mech. Phys. Solids, 42 (1), 1-20.
Budiansky, B. Hutchinson, J. W. and Ewans, A. C. (1986) Matrix fracture in fiber reinforced ceramics. J.
mech. Phys. Solids, 34, 167-189.
Carter, W. C., Butler, E. P. and Fuller, E. R. Jr., (1991) Micro-mechanical aspects of asperity-controlled
friction in fiber-toughened ceramic composites. Scripta Met. 25, 579-584.
Dugdale, D. S., (1960) Yielding of steel sheets containing slits. Journal of the Mechanics and Physics of
Solids, 8, 100-104.
Erdogan, F. and Joseph, P. F. (1989) Toughening of ceramics through crack bridging by ductile particles. J.
Am. Ceram. Soc. 72 (2), 262-270.
Marshall, D. B. and Cox, B. N. (1987) Tensile fracture of brittle matrix composites: influence of fiber strength.
Acta Met., 35, 2607-2619.
Marshal, P., and Price, J., (1991) Fibre/matrix interface property determination. Composites, 22 (1), 53-57.
Meda, G. and Steif, P. S. (1994a) A detailed analysis of cracks bridged by fibers - I. Limiting cases of short
and long cracks. J. Mech. Phys. Solids. 42 (8), 1293-1321.
Meda, G. and Steif, P. S. (1994b) A detailed analysis of cracks bridged by fibers - II. Cracks of intermediate
size. J. Mech. Phys. Solids. 42 (8), 1323-1341.
Muskhelishvili, N. I.., (1963) Some basic problems of the theory of elasticity. Noordhoff, Groningen, Holland.
Muskhelishvili, N. I., (1972) Singular Integral Equations. Noordhoff, Groningen, Holland.
Mumm, D. R. and Faber, K. T., (1995) Interfacial debonding and sliding in brittle matrix composites
measured using an improved fiber pullout technique. Acta Metall. Mat. 43 (3), 1259-1270.
Pagano, N. J. and Dharani, L. R. (1990) Micromechanical models for brittle matrix composites. Fiber
reinforced ceramic composites. Materials, processing and technology. (K. S. Mazdiyasni, ed.) pp. 40-
62.
Rose, L. R. F. (1987) Crack reinforcement by distributed springs. J. mech. Phys. Solids 35, 383-405.
Rubinstein, A. A., (1985) Macrocrack interaction with semi-infinite microcrack array, International Journal of
Fracture, 28, 113-119.
Rubinstein, A. A., (1986) Macrocrack - microdefect interaction. J. Applied Mechanics, 53, 505-510.
Rubinstein, A. A., (1987) Semi-infinite array of cracks in a uniform stress field. Engineering Fracture
Mechanics, 26 (1), 15-21.
Rubinstein, A. A., (1993) Micromechanical analysis of the failure process in ceramic matrix composites.
Journal of Engineering for Gas Turbines and Power, Transactions of the ASME, 115, 122-126.
Rubinstein, A. A., (1994) Strength of fiber reinforced ceramics on the basis of a micromechanical analysis. J.
Mech. Phys. Solids. 42 (3), 401-422.
384 A. A. RUBINSTEIN

Rubinstein, A. A., (1997) Micromechanical Approach to Failure Process in Composites, Advances in Fracture
Research. (B. L. Karihaloo, Y-W. May, M. I. Ripley and R. O. Ritchie, Editors) ICF9 Sydney,
Australia,. 2, 631- 642. Pergamon.
Rubinstein, A. A., (1998) Fracture Analysis of Composites by a Micromechanical Approach. Composites
Science and Technology, 58, 1785 - 1792.
Rubinstein, A. A. and Xu, K. (1992) Micromechanical model of crack growth in fiber reinforced ceramics. J.
Mech. Phys. Solids. 40 (1), 105-125.
Rubinstein, A. A., and Wang, P., (1996) Failure development in particulate composite, Durability and
Damage Tolerance of Composite Materials. AD-Vol. 51/MD-Vol. 73, Proceeding of the ASME
Aerospace and Materials Divisions, (Editors: W. S. Cham, M. L. Dunn, W. F. Jones, G. M. Newas, P.
V. D. McLaughlin and R. C. Wethehold) Book No. G01026-1996, ASME, 415-425.
Rubinstein, A. A., and Wang, P. (1998a) The fracture toughness of particulate-reinforced brittle matrix, J.
Mech. Phys. Solids. 46 (7), 1139-1159.
Rubinstein, A. A., and Wang, P. (1998b) Micromechanics of failure in metal matrix composites. Transactions
of the CSME, .22 (4B). 457-465.
Sigl, L. S., Mataga, P. A., Dalgleish, B. J., McMeeking, R. M., and Evans A. G. (1988) On the toughness of
brittle materials reinforced with a ductile phase. Avta Metall. 36 (4), 945-953.
Tvergard, V., (1992) Effect of ductile particle debonding during crack bridging in ceramics. In. J. Mech. Sci.
34 (8), 635-649.
Tvergard, V., (1995) On the micromechanics and fracture of ceramics. Fracture of Brittle Disordered
Materials: Concrete, Rock and ceramics. Edited by G. Baker and B. L. Karhaloo, F & FN Spon,
London, UK, 361-375.
Venkateswara Rao, K. T., Soboyejo, W. O., and Ritchie, R. O., (1992) Ductile-phase toughening and fatigue-
crack growth in Nb-reinforced molybdenum disilicide intermetallic composites. Metallurgical
Transactions A, 23A, 2249-2257.
Wang, P., (1998) Micromechanical Analysis of Failure Development in a Class of Composite Materials, Ph.
D. Dissertation, Tulane University.
J-INTEGRAL APPLICATIONS TO CHARACTERIZATION AND
TAILORING OF CEMENTITIOUS MATERIALS

VICTOR C. LI
Department of Civil and Environmental Engineering,
University of Michigan
Ann Arbor, MI, 48109-2125

1. Introduction

The J-integral introduced by Professor James R. Rice (Rice, 1968) has found extensive
applications in a broad variety of engineering materials. In the last decade, fracture
characterization of concrete and deliberate tailoring of fiber reinforced cementitious
composites have made great strides. The J-integral has played an important role in these
advances. This article reviews the contributions of the J-integral in three distinct but
related areas: (a) Characterization of the fracture process of cementitious materials, (b)
Testing methodology for the tension-softening constitutive relation in cementitious
materials, and (c) Design of cementitious composites with ultra high ductility.
Cementitious materials include a broad variety of engineering materials mostly
used in civil engineering applications. These include the ubiquitous concrete made of a
composition of aggregates with cement as binder. When the stone aggregates are
replaced by sand particles, the composite is referred to as a mortar. Fiber reinforced
concrete (FRC) is concrete containing a small amount of fiber, typically less than a few
percent by volume, and usually in discontinuous form. In recent years, the trend in high
performance cementitious composites has been in the use of mortar as the matrix
reinforced with an increasingly broad choice of fiber types. In the early days, high
performance cementitious composites are synonymous with high fiber content
composites. With improved understanding of the micromechanisms responsible for
multiple cracking and pseudo strain-hardening, some of these high performance
cementitious composites can now be engineered with only two percent or less of fibers,
making them viable economically and processing-wise for use in large scale structural
applications. In all of these cementitious materials, a common theme is that the
aggregates, sand particles or fibers serve as bridging elements when cracks traverse the
cement matrix. The fracture process and mode of failure are strongly influenced by the
properties of the bridging tractions working against crack opening and extension.
The non-linear fracture process in concrete is widely accepted in the engineering
community, and more accurate prediction of fracture load in concrete elements is now
possible. There is a gradually expanding, although still somewhat limited, adoption of
fracture based safe design of concrete structural elements. The application of high
performance fiber reinforced cementitious composites in load carrying structures is
emerging. A rapid growth in this area is expected in the next few years, especially in
Japan.
385
T.-J. Chuang and J. W. Rudnicki (eds.),
Multiscale Deformation and Fracture in Materials and Structures, 385406.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
386 V. C. LI

A detailed account of the fracture processes in cementitious materials can be found


in Li and Maalej (1996a,b). The early use of the J-integral for toughness characterization
of concrete was proposed by Halvorsen (1980) and Mindess et al (1977).
This article is written in honor of Professor James R. Rice, on the occasion of
celebrating his 60 t hbirthday.

2. Fracture Models for Concrete and Fiber Reinforced Cementitious


Composites

Hillerborg (1976, 1983) was one of the first to recognize the importance of aggregate
bridging in concrete and fiber bridging in FRC in the fracture processes of these
material. By including the often large scale process zone as an extension of the traction
free crack, the word fictitious crack was coined. In order to predict or simulate crack
propagation in these materials, fracture models are needed.
Whatever the source of crack face traction is, it is convenient to consider the
process zone as the near tip crack segment containing a line of springs tying the crack
faces. While any spring will invariably resist crack opening, the amount of energy
absorption and many macroscopic fracture behavior will depend on the detail behavior
of these springs. In general, the springs can be linear or non-linear, hardening or
softening. In the case of softening, it can be a result of the same physical process
leading to crack tip extension and therefore the presence of the process zone implies
cancellation of the crack tip singularity. We refer to such process zone as having
cohesive behavior. It is also possible to have the springs and crack tip extension as a
result of distinctly different physical processes, as in the case of a fiber reinforced
cement. In this case, the spring actions are associated with fiber bridging, whereas the
crack tip extension is a result of breaking down of the cement material. We refer to such
process zone as having bridging behavior. For a bridged crack, the presence of the
bridging zone can co-exist with a crack tip singularity. Such distinction between a
cohesive crack and a bridged crack was first recognized by Cox and Marshall (1994).
Cohesive crack models have been considered in a variety of contexts. Barenblatt
(1962) assumed the cohesion on the crack faces to be provided by the forces resisting
the separation of the layers of atoms in metals. Rice (1980) studied rock break down at
ends of earthquake shear faults. Hillerborg (1983) considered aggregate and ligament
bridging in cracked concrete producing a tension-softening behavior. The bridged crack
model appears most appropriate for fiber reinforced cementitious materials.
Consider a crack with a process zone of arbitrary size (Fig. 1). Traction in the
process zone takes a general relationship between crack face traction versus crack
opening ( ). The spring law can exhibit hardening and softening, with spring force
falling to zero at a critical crack opening t*. A relationship between such a general
spring law and the crack driving force J can be derived by adopting the contours shown
in Figure 2, and invoking the path independent property of the J-integral.
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 387

Figure 1. Line-Spring Concept of Process Zone Governing Crack Growth

Figure 2. J-integral Contours Chosen for Process Zone Analyses

The result for Mode I is

(1)

where is the relative crack opening displacement.


We consider three crack models corresponding to different crack tip and spring
behavior.

2.1 COHESIVE CRACK MODEL

When the presence of the cohesive zone is a direct result of the crack tip break down
process, as is often assumed to be the case in concrete, the crack tip singularity must
vanish, i.e.

(2)

Equation (1) then implies


388 V.C.LI

(3)

Figure 3 illustrates the tension-softening stress profile (x) in the process zone. Note
that the stress profile is continuous as it makes the transition from intact material ahead
of the fictitious crack tip ( =0) to the tension-softening material behind the crack tip.

Figure 3. (a) Stress Profile (x) in Cohesive Crack Model and (b) Corresponding ( ) Relationship. Note
Crack Opening 1 at Physical Crack Tip

The integral above can be re-written as

(4)

if Q is taken outside the process zone, since for > t, = 0 so that the integration for
> t can be truncated. Because the traction free crack must propagate when the crack
mouth opening exceeds t *, Eqn. (4) affords a definition of the critical J value, or

(5)

Equation (5) denotes an upper limit of J with no restriction on the size of the cohesive
zone. It corresponds to the critical value of non-linear fracture parameter J when
traction free crack extension initiates. Hence

J = Jc (6)
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 389

with Jc defined in (5) can be considered a crack initiation condition in large scale
cohesive zone crack model.
For the special case of small scale yielding, i.e. if the process zone is small
compared to all other characteristic dimensions in the problem, then J Q G, the
standard energy release rate crack driving force parameter. Further, at imminent
propagation ( t t*), G G c. This affords a physical interpretation of Gc in terms
of the inelastic behavior of the process zone material, i.e.,

(7)

or graphically, Gc represents the area under the spring law (Fig. 3b).

2.2 BRIDGED CRACK MODEL

For the bridged crack model, the presence of the process zone does not cancel the crack
tip singularity. This is the case of fiber reinforced cementitious composites, in which
the crack tip singularity can be associated with the fracture toughness of the cement,
while fiber bridging provides the spring tractions on the crack wake. Figure 4 illustrates
the bridging stress profile (x) in the process zone. Note that the stress profile is
discontinuous as it makes the transition from intact material ahead of the fictitious crack
tip to the bridging material behind the crack tip.

Figure 4. (a) Stress Profile (x) in Bridged Crack Model and (b) Corresponding () Relationship.

Assuming small scale yielding of the material ahead of the bridging zone, the contour p
can be shrunk onto the crack tip (but remains in the K-dominant zone, assuming that it
tip
exists), and we have J P G c , where G ctip denotes the toughness of the crack tip
material independent of the bridging zone processes. Equation (1) then becomes
390 V. C. LI

(8)

Again, since becomes zero when t exceeds t *, a critical value of J can be defined for
extension of the traction free crack:

(9)

For large scale bridging then, crack extension initiates when

J = Jc (10)

with Jc defined as in (9).


In the case of small bridging length compared to all other length scales in the
problem (small scale bridging), Jc and Gc coincides and

(11)

2.3 EMBEDDED PROCESS ZONE MODEL

The Bridged Crack Model discussed above assumes small scale yielding for the crack
tip material. However, this does not have to be the case. The recently developed highly
ductile Engineered Cementitious Composites (Li and Hashida, 1993) is a good example.
In such materials, the fiber bridging zone is embedded inside a volume of material
undergoing inelastic deformation (See also Figure 11 in Section 4). This suggests an
Embedded Process Zone Model shown schematically in Figure 5 together with a non-
linear stress-strain curve depicting the behavior of the inelastic zone (shaded area)
embedding the process zone.

Figure 5. (a) Embedded Process Zone Model, with Inelastic Behavior in Shaded Volume of Material
Represented by Non-Linear Curve in (b)

Equation (1) then gives


J-INTEGRAL FOR CEMENTITIOUS MATERIALS 391

(12)

where Jm represents the inelastic energy absorption inside the volumetrically distributed
inelastic zone. When both the off-crack-plane inelastic zone and the on-crack-plane
(cohesive) process zones are fully developed, the maximum value of J is reached. Thus

(13)

For most materials best described by the Bridged Crack Model, the crack tip toughness
is usually much smaller than the energy consumed in the bridging zone. For materials
that can be described by the Embedded Process Zone Model, experimental
measurements have indicated a Jm comparable in magnitude to the energy absorbed in
the process zone (second term on the right hand side of (13)) (Maalej et al, 1995a).
Again, for large scale process zone embedded inside an inelastic zone, the fracture
criterion in terms of J will be
J = Jc (14)
with Jc defined as in (13).
In the case of small process zone length and small inelastic off-crack-plane zone
compared to all other length scales in the problem, Jc and Gc coincides and

(15)

identical to the energy release rate of (11). In this limit, J m = G ctip.

3. J-Based Fracture Testing in Tension-Softening Material

It can be seen from the above discussion on fracture process characterization that the
() curve plays an important role as constitutive relation of the line-springs in the
fracture process zone. It is therefore important to have experimental methodology or
micromechanics based modeling to determine ( ). In the following, an experimental
technique (Li et al, 1987; Leung and Li, 1989; Hashida et al, 1993; Li et al, 1994) for
the determination of ( ) taking advantage of the J-integral is briefly reviewed.
Micromechanics based modeling of () for various fiber reinforced cementitious
materials can be found in Li and co-workers (1992, 1995, 1996, 1997).
The experimental technique to be discussed is based on the Compliance Test, first
used by Landes and Begley (1972) for elastic-plastic metallic materials, and utilizes the
interpretation of J as the difference in potential energy for a differential change in crack
length:

(16)
392 V.C.LI

where B is the specimen thickness. Because crack length change is often difficult to
measure, we use a pair of specimens identical in every respect except for a small
difference in crack length. The test can be carried out with any specimen geometry, and
the resulting value of J c should in principle be the same. The compact tension specimen
can be a convenient choice.
Suppose the two specimens have initial crack lengths of a1 and a2 , where a2 is
slightly larger than a1. For a valid test, a 2-a 1 should be smaller than all other dimensions
in the specimen, including the thickness B.
The load P and load point displacement are measured for each specimen (Fig.
6a). Due to process zone growth, the P curve can be nonlinear. Obviously the
specimen with longer crack will have larger displacement value for a given load level
(more compliant). For any fixed , the area between these two curves may be
interpreted as
(17a)
so that

(17b)

Since a1, a2 and B are known a priori , J can be calculated for each value of . This is
shown in Fig. 6b as a J curve. The plateau value of J is interpreted as the Jc value
associated with the full development of the process zone.
During specimen loading, the crack tip opening displacement is also monitored.
Thus a triplet of Load P, Load-point displacement , and crack tip displacement t is
recorded during the test, and the specimens are loaded to beyond the peak load into the
softening regime. Figure 6c shows a versus t correlation curve.

Figure 6a. P Record for an FRC Specimen (after Li and Ward, 1989)
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 393

Figure 6b: J Record

Figure 6c. t Record

The correlation between and t were used by Li and co-workers to deduce the
curve. By using the relationship between J and () (Equation (4)) for tension-softening
materials, the curve (Fig. 7) can be obtained by differentiation. That is,

(18)
394 V.C.LI

Figure 7. Deduced Tension-Softening Curve

Figure 7 also shows the tension-softening curve of the same FRC material obtained from
a uniaxial direct tension test. The good comparison of the two curves suggests the
accuracy of the J-based testing technique. The J-based method is particularly suitable
for brittle materials with sharp dropping curves for which the direct tension test
may be difficult to carry out due to load instability beyond peak. The J-based testing
technique was originally developed for mortar and concrete (Li et al, 1987; Teramura et
al, 1990), but has since been applied to fiber reinforced composites (Leung and Li,
1989; Rokugo et al, 1989; Li et al, 1994; Hashida et al, 1994), and rocks (Chong et al,
1989; Hashida, 1990). The J-based technique has been extended to relationship
determination for materials in which the crack tip singularity is not canceled (Li et al,
1994).

4. Steady State Cracking and Strain-Hardening Design

Although most cementitious materials are considered brittle (e.g. cement), or quasi-
brittle (e.g. concrete and FRC), it is possible to design cementitious composites with
extremely ductile behavior. One such material, known as Engineered Cementitious
Composite (ECC for short), exhibits tensile strain capacity up to 7.5% (Li et al, 1996).
The design of such materials is based on the J-integral analyses of steady state cracking
and the micromechanics of fiber bridging.
In order to achieve the desirable pseudo-strain hardening behavior, two criteria
must be satisfied (Li and Leung, 1992; Li et al, 1996): (i) steady state cracking criterion,
and (ii) first crack criterion, which requires the first cracking stress to be lower than the
maximum fiber bridging stress. Additional cracks (multiple cracking) can then form on
further loading.
The steady state criterion has been studied by a number of researchers, (see, e.g.
Marshall and Cox (1988); Li and Wu, (1992); and Li and Leung, (1992)). In fiber
composites, the extension of a matrix crack is accompanied by fiber bridging across the
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 395

crack flanks. As the matrix crack extends, the bridging zone increases in length. During
crack opening, the bridging stress increases as fiber/matrix interfaces debond and the
debonded segments of fibers stretch (hardening spring behavior). When the bridging
stress increases to the magnitude of the applied load, the crack flanks flatten to maintain
the constant applied stress level (Li and Wu, 1992). This load level is termed the steady
state cracking stress ss .
Based on a J-integral analysis of a steady state crack, Marshall and Cox (1998)
showed that

(19)

where J tip refers to the crack tip toughness. In most fiber reinforced cementitious
composites with less than 5% fiber volume fraction, J tip can be approximated as the
cementitious matrix toughness. The steady state stress ss and the flattened crack
opening ss are related via the bridging law ( ). The right hand side of (19) is known
as the complementary energy of fiber bridging, and corresponds to the shaded area
above the ( ) curve in Fig. 8. For steady state cracking to occur at all, the steady state
cracking stress must be less than the maximum bridging stress 0 in the bridging law.
That is,
(20)
Equation (19) and (20) together imply

(21)

Figure 8. Complementary Energy Concept of Fiber Bridging


396 V. C. LI

Equation (21) provides a general condition for transition from quasi-brittle to strain-
hardening failure mode, and highlights the importance of the total complementary
energy (right hand side of (21)) in composite design.
For Eqn. (21) to be useful in fiber, matrix and interface tailoring, it will be
necessary to determine the bridging law () specific for a given composite system. In
fiber reinforced cementitious composites in which the fibers are randomly oriented and
in which pull-out (rather than fiber rupture) are expected, the bridging laws developed
by Li and Leung (1992) can be summarized in the following form:

(22)

where is the crack opening corresponding to the maximum


bridging stress

(23)

Corresponding equations for cases where fibers can rupture and for cases where fibers
are of variable diameters can be found in Maalej et al (1995b), and Li and Obla (1996).
In Eqs. (22) and (23), V, L, d, and E f are the fiber volume fraction, length, diameter
and Youngs Modulus, respectively. is the fiber/matrix interface friction, and the
snubbing factor

(24)

where f is a snubbing coefficient which must be determined experimentally for a given


fiber/matrix system (Li et al, 1990). The snubbing coefficient raises the bridging stress
of fibers bridging at an angle inclined to the matrix crack plane, appropriate for flexible
fibers exiting the matrix analogous to a rope passing over a friction pulley. Finally, =
(VE )/(VmEm), where V m and Em are the matrix volume fraction and Youngs Modulus,
respectively.
The condition for steady state cracking expressed in Eqn. (21) can now be
interpreted as a critical fiber volume fraction above which the composite will show
pseudo strain-hardening. Using (22) in (21), this critical fiber volume fraction can be
defined in terms of the fiber, matrix and interface parameters (Li and Wu, 1992):
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 397

(25)

Equation (25) is important for composite design. It provides guidelines for tailoring the
crit
microparameters such that V is minimized (Li, 1998). Strain-hardening composites
can then be designed with the minimum fiber content.

Figure 9. Uniaxial Tensile Stress-Strain Curve of an ECC with CO2 gas plasma treated PE fibers

In what follows, we describe the mechanical properties for an ECC. Unless otherwise
stated, the ECC referred to contains two volume percent of polyethylene fibers. Using
Eq. (25) and appropriate parametric values (see Li, 1998), the critical fiber volume
fraction is estimated to range between 0.5% and 1%. Hence a composite with 2% fiber
should satisfy the condition of pseudo strain-hardening, and exhibit high strain capacity
after first cracking.
Figure 9 shows the stress-strain curves from uniaxial tension tests. The ECC strain
hardens to an average strain at peak stress cu approximately equal to 5.6 % (about 560
times the strain capacity of the unreinforced matrix). For this composite, real-time
observation showed that multiple cracking occurred with many sub-parallel cracks
across the specimen during strain-hardening. Beyond peak stress, localized crack
extension occurred accompanied by fiber bridging. The multiple cracking pattern of a
specimen is shown in Figure 10.
398 V. C. LI

Figure 10. Multiple Crack Pattern of an ECC at About 4.2% Tensile Strain

The total fracture energy of ECC was determined (Li and Hashida, 1993; Maalej et
al, 1995) by means of the J-based technique (Eqn. 17b) and using a set of DCB
specimens with different notch lengths. Concurrently with the tests, damage evolution
on the specimen surface was recorded using a camera.
Figure 11 presents the damage evolution recorded at various loading stages. It is
particularly noted that an extensive microcrack damage zone spreads around the notch
tip before the localized crack starts to grow. Significant energy absorption is therefore
expected from the off-crack-plane volumetric inelastic deformation process. The total
fracture energy measured for this ECC was 27 kJ/m2 ,with approximately over half of
this energy consumed in the inelastic damage process occupying an area of 1150 cm 2
around the crack tip, and the rest coming from the pull-out of fibers on the crack wake.
For the ECC, the Embedded Process Zone Model discussed in Section 2.3 is most
appropriate in describing its fracture process.
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 399

Figure 11. DCB Specimen Showing Evolution of Notch Tip Inelastic Zone at Four Loading Stages.

The notch-sensitivity of ECC has been examined with double edge notched specimens.
Test results are shown in Figure 12, which plots the peak load as a function of the
reduced section of the notched specimens. The data of the notched specimen lying near
(and actually slightly above) the linear line suggests that these composites are notch-
insensitive. The surface of the notched specimen (Fig. 13) shows multiple cracks
typical of strain hardening fiber reinforced composites. Although the ultimate localized
fracture is in the reduced section, multiple cracking spreads along the full length of the
specimens prior to final failure. These results suggest that highly damage tolerant
structural behavior can be achieved.
400 V. C. LI

Figure 12. Nominal Failure Load of Double Edged Notch Specimen

Figure 13. Damage Pattern of DEN Specimen

The ability of ECC to deform non-linearly with strain-hardening in tension combined


with high damage tolerance suggests its use in concrete elements which require bolt
jointing. An experimental study (Li and Kanda, 1998) was carried out to determine the
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 401

response of an ECC slab under circular indentor load (Fig. 14). For this ECC material,
1.25% of PVA fiber was employed. In uniaxial tension the strain capacity was 5%. As
control, a similar slab with plain mortar was tested under the same load configuration.
Three different size indentors were used. Figure 15 shows the load-deformation
(indentor deflection) curves for the ECC and the mortar specimens (used as control).
Each specimen type was loaded with three bearing sizes expressed as a percentage of
slab surface area. While the load capacity in each case is comparable, it is clear that the
deformation capacity of the ECC slab is about one order of magnitude higher than that
of the mortar slab at failure.
Figure 16 shows the failure pattern of a mortar specimen, which fractures brittlely
into several pieces as expected. The corresponding failure of the ECC specimen is
much more ductile. Even as the indentor penetrates the surface of the slab, the
surrounding material undergoes inelastic damage with no fractures (Figure 17a). Figure
17b gives an enlarged view of the indentor punch.
The results of this test confirms the notion that the strain-hardening and damage
tolerance of ECC can be very effective in alleviating the high stress concentrations
experienced by structural elements whenever steel and concrete materials come into
contact with each other. Such elements may include concrete embedded steel anchors,
and connections in hybrid concrete/steel structural members.

Figure 14. Geometry of indentor and ECC/Mortar Slab

Figure 15. Load -Deformation Curves for (a) Mortar, and (b) ECC Slab
402 V. C. LI

Figure 16. Fracture Failure of A Mortar Specimen

Figure 17. (a) Ductile Indented Pattern of A ECC Specimen, and (b) Enlarged View

ECC is now being investigated for structural applications (Li and Kanda, 1998;
Fukuyama et al, 1999; Parra-Montesinos and Wight, 2000). A recent study (Fischer and
Li, 2000) on exploiting the strain-hardening and multiple cracking behavior of ECC in
highly earthquake resistant building systems involves fully reversed cyclically loaded
flexural members. These flexural members are made of ECC reinforced with
longitudinal steel or FRP rods. Figure 18 shows the deformed shape of an ECC/aramid-
FRP element at 10% interstory drift. Microcracks less than 200m are formed along the
full length of the element. The corresponding hysteretic loops indicating large
deformation capacity but with low residual deformation (especially for drift less than
5%) are also shown. In contrast to typical concrete/steel element behavior, no spalling
of the ECC matrix or buckling of the axial reinforcements are observed. Corresponding
tests with ECC/steel elements show extremely high energy absorption behavior (Fischer
and Li, 2000). These characteristics can be used to design building systems with high
safety as well as minimal post-earthquake repair needs.
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 403

Figure 18. (a) FRP Reinforced ECC Flexural Element at 10% drift and (b) load deformation behavior
404 V. C. LI

5 . Conclusions and Further Discussions

The J-integral has played an important role in the understanding of the mechanics and
the engineering of the composites of cementitious materials. This article reviews the use
of the J-integral in fracture process characterization, tension-softening curve
determination, and microstructural tailoring of cementitious materials. Advancements
in these areas provide important tools for failure analyses of structures of cementitious
materials, as well as tools for cementitious composite design for safe structures.
Before closing, we note an often misdirected criticism of the application of the J-
integral to tension-softening materials. The criticism is directed at the fact that since the
J-integral rests on the assumption of non-linear elasticity, and since the process zone
material softens inelastically, the application of the J-integral to tension-softening
materials cannot be valid. The apparent paradox is resolved if it is understood that the J-
integral contour (see, e.g. the contours QP+ and QP- in Fig. 2) is placed in the elastic
material adjacent to a line of springs (representing the softening material) which can
unload. Then the unloading of the springs (softening branch of the curve) during
crack opening causes a corresponding elastic unloading of the material in which the
contour is placed. This can be better envisioned with a tensile specimen of a quasi-
brittle material. Once a localized fracture zone is formed, the material in the fracture
zone unloads (inelastically), while to maintain equilibrium, the material outside the
fracture zone, which remains elastic also unloads, but unloads elastically. This
phenomenon was nicely illustrated with concrete specimens in uniaxial tensile
experiments carried out by Petersson (1981). The strain/displacement gage across the
fracture zone shows inelastic unloading (stress drop with increase in crack opening), but
the gages outside the fracture zone unloads elastically (decreasing stress with decreasing
strain deformation retracing the elastic loading). Hence for the application of the J-
integral to tension-softening materials, there is no violation of the requirement of the J-
integral as long as the unloading of the material outside the line-spring is elastic.

References

Barenblatt, G. I. (1962) The mathematical theory of equilibrium cracks in brittle fracture, Advanced Applied
Mechanics 7, 55-125.
Chong, K.P., Li, V.C., and Einstein, H.H. (1989) Size effects, process zone, and tension softening behavior in
fracture of geomaterials, International J. of Engineering Fracture Mechanics 34(3), 669-678.
Cox, B. and Marshall, D. (1994) Concepts in the fracture and fatigue of bridged cracks, Overview No 111.
Acta Meta. Mate. 42, 341-363.
Fischer, G. and Li, V.C. (2000) Structural composites with ECC, to appear in the Proceedings of the ASCCS-
2000.
Fukuyama, H., Y. Matsuzaki, K. Nakano, and Y. Sato (1999) Structural Performance of Beam Elements with
PVA-ECC, in H. Reinhardt and A. Naaman (eds.), Proc. of High Performance Fiber Reinforced Cement
Composites 3 (HPFRCC 3), Chapman & Hall, pp. 53 l-542.
Halvorsen, G.T. (1980) J-integral study of steel fiber reinforced concrete, International J. Cement Composites,
2(1) 13-22.
Hashida, T. (1990) Evaluation of fracture processes in granite based on the tension-softening model, In S.P.
Shah, S.E. Swartz & M.L. Wang (eds), Micromechanics of Failure of Quasi-Brittle Materials, Elsevier
Applied Science, London, 233-243.
Hashida, T., Li, V.C., and Takahashi, H. (1994) New development of the J-based fracture testing technique for
ceramic matrix composites, J. American Ceramic Society 77(6), 1553-1561.
J-INTEGRAL FOR CEMENTITIOUS MATERIALS 405

Hillerborg, A. (1983) Analysis of One Single Crack, in F.H. Wittmann (ed.) Fracture Mechanics of Concrete,
Elsevier Science Publisher, B.V., Amsterdam, pp. 223-250.
Hillerborg, A., Modeer, M., and Petersson, P. E. (1976) Analysis of crack formation and crack growth in
concrete by means of fracture mechanics and finite elements. Cement and Concrete Research 6, 773-
782.
Landes, J.D. and Begley, J.A. (1972) The Effect of Specimen Geometry on J I C, in Stress Analysis and Growth
of Cracks, ASTM STP 514, ASTM, Philadelphia
Leung, C.K.Y., and Li, V.C. (1989) Determination of fracture toughness parameter of quasi-brittle materials
with laboratory-size specimens, J. Materials Science 24, 854-862.
Li, V.C. (1992) Post-crack scaling relations for fiber reinforced cementitious composites, ASCE J. of
Materials in Civil Engineering, 4(1), 41-57.
Li, V.C. (1998) Engineered Cementitious Composites Tailored Composites Through Micromechanical
Modeling, in N. Banthia, A. Bentur, A. and A. Mufti (eds.) Fiber Reinforced Concrete: Present and the
Future, Canadian Society for Civil Engineering, Montreal, pp. 64-97.
Li, V.C., Chan, C.M., and Leung, C.K.Y. (1987) Experimental determination of the tension-softening curve in
cementitious composites, J. Cement and Concrete Research 17-3, 441-452.
Li, V.C. and Leung, C.K.Y. (1992) Steady state and multiple cracking of short random fiber composites,
ASCE J. of Engineering Mechanics 118(11), 2246-2264.
Li, V.C. and Hashida, T. (1993) Engineering ductile fracture in brittle matrix composites, J. of Materials
Science Letters 12, 898-901.
Li, V.C. and Kanda, T. (1998) Engineered cementitious composites for structural applications, ASCE J.
Materials in Civil Engineering 10(2), 66-69.
Li, V.C. and Maalej, M. (1996a) Toughening in cement based composites, Part I: Cement, mortar and
concrete, J. of Cement and Concrete Composites 18(4), 223-237.
Li, V.C. and Maalej, M. (1996b) Toughening in cement based composites, Part II: Fiber reinforced
cementitious composites, J. of Cement and Concrete Composites 18(4), 239-249.
Li, V.C., Maalej, M., and Hashida, T. (1994) Experimental determination of stress-crack opening relation in
fiber cementitious composites with crack tip singularity, J. Materials Science 29, 2719 - 2724.
Li, V.C., Mihashi, H., Wu, H.C., Alwan, J., Brincker, R., Horii, H., Leung, C., Maalej, M., and Stang, H.
(1996) Micromechanical models of mechanical response of HPFRCC, in A.E. Naaman and H.W.
Reinhardt (Eds.) High Performance Fiber Reinforced Cementitious Composites, RILEM Proceedings 31,
pp. 43-100.
Li, V.C. and Obla, K. (1996) Effect of fiber diameter variation on properties of cement based matrix fiber
reinforced composites, Composites Engineering International Journal Part B 27B, 275-284.
Li, V.C. and Ward, R. (1989) A novel testing technique for post-peak tensile behavior of cementitious
materials, in H. Mihashi, H. Takahashi, and F.H. Wittmann (eds.) Fracture Toughness and Fracture
Energy Test Methods for Concrete and Rock, Balkema, Rotterdam, pp. 183 195.
Li, V.C., Wang, Y., and Backer, S. (1990) Effect of inclining angle, bundling, and surface treatment on
synthetic fiber pull-out from a cement matrix, J. Composites 21(2), 132-140.
Li, V.C. and Wu, H.C. (1992) Conditions for Pseudo strain-hardening in fiber reinforced brittle matrix
composites, J. Applied Mechanics Review 45(8), 390-398.
Li, V.C., Wu, H.C., and Chan, Y.W. (1996) Effect of plasma treatment of polyethylene fibers on interface and
cementitious composite properties, J. of American Ceramics Society 79(3), 700-704.
Lin, Z. and Li, V.C. (1997) Crack bridging in fiber reinforced cementitious composites with slip-hardening
interfaces, J. Mechanics and Physics of Solids 45(5), 763-787.
Maalej, M., Hashida, T., and Li, V.C. (1995a) Effect of fiber volume fraction on the off-crack-plane fracture
energy in strain-hardening engineered cementitious composites, J. Amer. Ceramics Soc. 78(12), 3369-
3375.
Maalej, M., Li, V.C., and Hashida, T. (1995b) Effect of fiber rupture on tensile properties of short fiber
composites, ASCE J. Engineering Mechanics 121(8), 903-913.
Marshall, D.B. and Cox, B.N. (1988) A J-integral method for calculating steady-state Matrix Cracking
Stresses in Composites, Mechanics of Materials 7, 127-133.
Mindess, S., Lawrence, Jr., F.V. and Kesler, C.E. (1977) The J-integral as a fracture criterion for fiber
reinforced concrete, Cement and Concrete Research, 7, 731-742.
Parra-Montesinos, G.J., and J.K. Wight (2000) Behavior and Strength of RC Column-to-Steel Beam
Connections Subjected to Seismic Loading, to appear in the Proceedings of the 12th WCEE.
Petersson, P-E, (1981) Crack Growth and Development of Fracture Zones in Plain Concrete and Similar
Materials, Report TVBM-1006, Lund, Sweden, 174pp.
406 V. C. LI
Rice, J. R. (1968) A path-independent integral and the approximate analysis of strain concentration by notches
and cracks. J. Applied Mechanics 35, 379.
Rice, J.R. (1980) The mechanics of earthquake rupture, in A.M. Dziewonski and E. Boschi (eds.) Physics of
the Earths Interior, Italian Physical Society/North Holland, Amsterdam.
Rokugo, K., Iwasa, M., Seko, S., and Koyanagi, W. (1989) Tension-softening diagams of steel fiher
reinforced concrete. In S.P. Shah, S.E. Swartz & B. Barr (eds.), Fracture of Concrete and Rock, Recent
Developments, pp. 513-522.
Teramura, S., Normura, N., Hashida, T., and Mihashi, H. (1990) Development of a core-based testing method
for determiningfracture energy and tension-softening relation of concrete, in S.P. Shah, S.E. Swartz and
M. L. Wang (eds.), Micromechanics of Failure of Quasi-Brittle Materials, Elsevier Applied Science,
London, pp. 463-473.
Author Index

Abouaf, M. 50 Betegon, C. 238


Abraham, F. 278 Biot, M.A. 108
Agrait, N. 82 Birnbaum, H. K. 31
Aldinger, F. 358 Biwa, S. 71,74
Alwan, J. 391 Blackburn, W.S. 206
Amazigo, J.C. 361,367 Bloyer, D.R. 370,379
Anderson P. M. 31,41,87,89 Blug, B. 49
Antonellini, M.A.161 Blunt, R. J. 367
Argon, A. S. 10,90 Bockris, J. OM. 43
Arzt, E. 50 Boniface, V. 183
Asaro, R.J. 184 Bordia, R.K. 50
Ashby, M.F. 50,53,61,65,367 Botsis, J. 363
Askenazi, D. 184,190 Bouvard, D. 50
Astiz, M.A. 217 Bowden, F. P. 82
Atkins, A.G. 223,225 Bower, A.F. 108,331
Aveston, J. 361 Boyce, M. C. 19
Boyle, R.W. 225
Bader, S.D. 108 Brinker, R. 391
Baker, S. 396 Brodbeck, D. 278
Ball, P. 109 Budiansky, B. 211,361,367
Banister, M. 367 Bullough, R. 112
Banks-Sills, L. 183-184, 188, 190 Butler, E.P. 366
Bao, G. 362,367 Byerlee, J.D. 164,177
Barenblatt, G.I. 386
Barlat, F. 17-21,26 Cahn, J.W. 108,112
Barnett, S.A. 88 Cammarata, R.C. 112
Bartelt, M.C. 108 Cao, H.C. 358
Barthel, E. 72 Carlsson, A.G. 238
Basinski, Z.S. 319,327 Carter, W.C. 366,108
Basinski, S.J. 319,327 Chambreuil, A. 31
Bassani, J. 184 Chan, C.M. 394
Baud, P. 170 Charalambides,P.G. 184
Baudin, P. 50 Charles, R. J. 332-333,335
Bayle, B. 31 Chen, B. 17
Beck, W. 43 Chen, X.F. 43
Begley, J .A. 206,219,391 Chen, C.H. 168
Beltz, G. E. 41,87,90,96,103,237, Chen, X. 31,44
239 Chen, Z. 223,225-227
Bernstein, I. M. 31,44 Chen, L.Q. 109,114,117
408 AUTHOR INDEX

Chen, X.-Y. 243,261-263, 266,271 Du, Z.Z. 50


Cheng, Y. 89 Duesbery, M.S. 90,93
Chenot, J.L. 50 Dugdale, D.S. 371
Cherepanov, G.P. 350,353 Dundurs, J. 89,186
Chien, W. Y. 1,3,8,9 Dutton, R. 338
Chitaley, A.D. 243,256
Cho, J.W. 311-312
Chong, K.P. 394 Einstein, H.H. 394
Chu, C. -C. 10 Eisele, U. 50
Chuang, T.-J. 51,107,112,331-333, Eliasi, R. 184,190
339-343, Elices, M. 217
Cleary, M.P. 164,177-178 Erdogan, F. 362,367
Clememens, B.M.88 Ernst, H.A. 226-227
Cocks, A.C.F 50,81,108 Eshelby, E.J. 123
Collier, L. 358 Evans, A.G. 346,358,361-362,367
Comsa, G. 108
Cook, W.R. 359 Faber, K.T. 366
Cooper, G.A. 361 Ferreira P. J. 31
Corten, H.T. 188 Fineberg, J. 277,290,295,299-300
Cotterell, B. 223,225,238 Finno,R.J. 169-170
Coussy, O. 171 Fischer, L.L. 237,241
Cox, B. 386,394 Fischer, G. 402
Cox, B. N. 362 Fleck, N. A. 72,81,238
Craft, D. 108 Foecke, T. 43
Cui, Y.L. 361 Freiman, S.W. 332
Cuitino, A.M. 312,327 Frenkel, J. 90,239
Curtin, W.A. 331 Freund, L. B. 89,108-109,112,289
Fuhirott, H. 324
Dalgleish, B.J. 362 Fukai, Y. 43
David, C. 161 Fukuyama, H. 402
Dawson, P.R 50 Fukuyama, S. 31
de la Figuera, J. 108 Fuller, Jr. E.R 332-333,339-343
Deng, X. 186
Desrues, J. 161,168 Gaebel, R. 50
Dharani, L.R. 362 Galvez, V.S. 217
Diercks, V. 108 Ganti, S. 244
Dimaggio, F.L. 161,169 Gao, H. 108,275,277-280,289-
Dimiduk,D.M. 88 290,296
Dolbow, J. 184 Garagash, D. 168
Doremus, P. 50 Gehrke, E. 332
Drory, M.D. 184 George, J. 108
Drugan, W.J. 243-244,246-247, Gerberich W.W. 31,43-44
256,261-263,266,271 German, M.D. 225
AUTHOR INDEX 409

Gill, S.P.A. 108 Huang, Y. 17


Gillia, O. 50 Huang, H. -M. 1,10-11,13-14
Glas, F. 109 Hui, C.-Y. 362,367
Gong, X. 109 Hull, D. 107
Gosz, M. 184 Hurtado, J.A. 89
Graf, A. 2 Hutchinson, J.W. 10,72,81,183,186,
Greenspan, D.C. 332 224-225, 238
Greenwood, J. A 72,74 Hwang, R.Q. 108
Greskovich, C. 65
Gross, S. 277,290,295,299-300 Igarashi, M. 31,44
Gurland, J. 9 Im, J. 10
Gurson, A. L. l-10, 17-29,277 Imbault, D. 50
Gurtin, M.E. 112 Irwin, G.R. 223,300
Guyer, J.E. 109 Issen, K.A. 161
Iwasa, M. 394
Hack, J. E 43
Hahnert, M. 332 Jackson, P. J. 327
Hall, E.O. 88 Jaffe, H. 359
Halvorsen, G.T. 386 Jaffe, B. 359
Han, C. 169 Jagota, A. 50
Hanada, S. 31 Jiang, J.S. 108
Hancock, J.W. 225,238,244 Johnson, H. 346
Harris, W.W. 169 Johnson, K. L 7 l-72.74,80-82,113
Hashida, T. 390-391,394,396,398 Joseph, P.F. 362,367
Haug, A. 349 Joyce, J.A. 225,233
Hazzledine, P.M. 88-89
Head, A.K. 89,98-99 Kachanov, M. 123,277
Hector,Jr.L.G. 152 Kagawa, K.I 51,107
Herbek, J. 108 Kanda, T. 400,402
Herring, C. 107 Kanters, J. 50
Hibbitt, D. H. 7 Kanwal, R.P. 123
Hill, R. 1,2,11,18,71-72,75,291 Karafillis, A. P. 18-19
Hillerborg, A. 386 Karapetian, E. 123
Hillert, M. 59 Katz, Y. 31,43-44
Hilliard, J.E. 108,112 Kawakami, T. 31
Hillig, W .B. 332-333,335 Kaxiras, E. 90,93
Hirth, J. P. 32-35,42, 89-90, 101, Keer, L.M. 123
105 Kelly, A. 361
Hoagland R. G 41 Kern, K. 108
Hom, C. L. 9 Kesler, C.E. 386
Horii, H. 391 Khachaturyan, A.G. 108-109
Hosford, W. F. 2,18 Kikuchi, M. 206
Hu, Z. 31 Kimura, H. 41
410 AUTHOR INDEX

Kitajima, S. 312 Magnin, T 31


Kiuchi, K 43 Majorana, C.E. 168,170
Klameth H.-K 31,44. Marder, M. 277,290,295,299-300
Klein, P.A. 275,279,,290,294,296 Marshal, P. 366
Knauss, W. 276 Marshall, D.B. 362,386,394
Knowles, J.K. 214 Martin, R.J.III 168
Koehler, J. 152-153 Mataga, P.A. 244,362
Koepke, B.G. 358 Matos, P.P.L. 184
Koike, S 43 Matsui, H 41
Koyanagi, W. 394 Matsuzaki, Y. 402
Kozak, V 50,55,56 Maugis, D 72
Krafft, J.M. 225 McBreen, J 43
Kraft, T 50 McClintock, F.A. 243,256
Krznowski, J.E. 89 McHenry, K.D. 358
Kubo, S. 205-206,217,219-220 McHugh, P.E 50
Kuhlmann-Wilsdorf, D. 152 McLean, D 37
Kuhn, L.T 50,55,65,108 McLellan, R. B 43
Kung, H. 88 McMahon, C. J 31
McMeeking, R.M 9,43,50,55,
Landes, J.D. 206,219,226-227,391 65,108,184,
Lanxner, M 31.44 225,349-51
Larsson, S.G. 238 Mear, M. E. 10
Larsson, R. 161 Meda, G. 362
Lawerence, Jr. F.V. 386 Merkle, J.G. 219
Leung, C.K.Y. 391,394,396 Mesarovic, S. Dj. 71-72
Levine, L.E. 145,147-149 Mihashi, H. 391
Li, J. 244 Mikeska, K.R 50
Li, J.C.M. 241 Mindess, S. 386
Li, D. 108 Miyamoto, H. 206
Li, V.C. 385-386,390-391 Miyata, K 31,44
Liao, K. -C. 2,3,13,19,25,27,29 Modeer, M. 386
Lii, M.J 31,43-44 Mohan, R. 312
Link, R.E. 225,233 Mokni, M 161,168
Lothe, J. 89-90,101,105 Mollema, P.N. 161
Lu, W. 107,109,115 Mooney, M.A. 169
Lufrano, J 43 Moran, B. 184
Lure, A.I. 123 Morgan, K. 337
Lynch, C.S. 358 Mori, K 50
Mullins, W.W. 107
Maalej, M. 386,391,394,396 Mumm, D.R. 366
MacEwen, S. R. 17 Mura, T. 87,89-90,96,98-99,103
Mader, S. 153 Murdoch, A.I. 112
Maeda, T. 220 Muskhelishvilli, N.I. 363,366,373
AUTHOR INDEX 411

Nabarro, F.R.N. 238 Pierce, D. 10


Nahta, R. 184 Plankensteiner, A.F 50,58
Nakamura, T. 184 Pohanka, R.C. 358
Nakano, K. 402 Pohl, K. 108
Namazue, H. 43 Pompe, W. 109
Needle, A. 2, 9-10, 18, 107, 278, Pond, Sr.R.B. 153
299 Ponte, C. P. 244
Neumann, P. 31,44,324 Price, J. 366
Ng, K.-O 109 Prigogine, I. 108
Niebus, H. 108
Nishimura, K. 220 Qu, J. 184
Noggle, T. 152-153
Nur, A 164, 177 Rafey, R. 278
Rahman, M. 123
Obla, K 394,396 Raisson, G 50
Ogbonna, N. 81 Rao, S.I. 88-89
Ogura, K. 205 Ravi-Chanda,K. 276
Ohji, K. 205206,217,219-220 Rayleigh, J.W.S.108
Ohta, M. 312 Reuter, W.G. 225
Olevsky, E.A 50 Rice, J.R. 11,18,31-35,37,39,
Olsson, W.A 161 41, 51,72,87,90, 96,
Ortiz, M. 279,312,327 103, 107, 112, 159,
ODowd, N.P. 225 182-183, 186, 205,
207, 211, 219, 223,
Pacheco, E.S. 87,89-90,96,98-99,103 225, 238-240, 244,
Pagano, N.J. 362 246-248, 311, 314,
Pak, Y.E. 358 317, 327, 346, 350,
Pan, J. 1,2,3,108 353-355, 357-358,
Paris, P. 2 19,224,226-227 362, 366-367, 370,
Paris, P.C. 319,327 376, 378-379, ~386,
Park, S.B. 351,354 394-398,400,402
Parks, D.M. 225,244 Rice, R.W. 358
Parra-Montesinos, G.J. 402 Riedel, H 49-50,54-58
Parteder, E 50,58 Rimmer, D.E. 107
Paterson, M.S. 163 Ritchie, R.O. 367,370,378-379
Pearson, J. 108 Robertson I. M 31
Pearson, J. M 81 Rodel, J 50
Peierls, R.E. 238 Rokugo, K. 394
Peraire, J. 337 Rose, L.R.F. 361
Peric, D. 160 Rosengren, G.F. 225
Petch, N. J. 88. Rosolowski, J.H 65
Petersson, P.E. 386,404 Rostek, A. 358
Phillips, R. 279 Rothe, W 31,44
412 AUTHOR INDEX

Royburd, A.L. 109 Stakgold, 1. 281


Rubinstein, A.A. 361-362,364-365 Stang, H. 391
Rubio, G 82 Stauffer, D. 147-148
Rudge, W. 278 Steif, P.S. 362
Rudnicki, J.W. 159,162,167-176 Sternberg, E. 214
Runesson, K. 160-162,171,178-179 Stingl, P 50
Storkers, B 71,73-75
Saeedvafa, M. 311 Sture, S. 160
Saje, M. 2,10 Suehiro, S. 217
Sanavia, L. 168 Sugimoto, H. 43
Sandler, I.S. 161,169 Sullivan, A.M. 225
Sato, Y. 402 Sumpter, A. 224
Schatz, A. 108 Sun, D.-Z. 50
Scherer, G.W 50 Sun, C.T. 35 1,354
Schneider, G.A. 358 Sun, Y. 41, 90,93
Schoeck, G. 37 Suo, Z. 107-109, 115, 184,
Schrefler, B.A. 168 238,358
Schubnel, A. 170 Suzuki T 43
Seah M. P 37 Svobada, J. 50,52-57
Seko, S. 394 Swinney, H. 277,290,295,299-300
Selvadurai, A.P.S. 123
Sevostianov, I. 123 Tabor, D. 82
Shafig, A.B. 363 Tadmor, E. 279
Sham, T.-L. 244 Takano, N. 31
Sharma, D.L. 123 Takasugi, T. 31
Shen, H. 43 Tang, Z. 331
Shen, J. 117 Tang, S. C. 1,3, 17
Sherman, D. 188 Taylor, J.E. 108
Shield, T.W. 311 Terasaki, F. 31
Shih, C. F. 184,225,312 Thompson A.W. 31,44
Shilkrot, L.E. 89 Thomson, R. M. 37,145,147-149
Shim, Y. 145,147 Timothy, S. P. 81
Shinagawa, K 50 Ting, T.C.T. 184,187
Shinn, M. 88 Tobin, A.G. 358
Sieradzki, K. 112 Tonda, M. 312
Sigl, L.S. 362 Tong, W. 152
Sih, G.C. 319,327 Toribio, J. 43
Sills, L.B 51,107 Tracey, D. M. 18,244
Simmons, D.C. 332 Travitzky, N. 184,190
Soboyejo, W.O. 367,378 Triantafyllidis, N.2
Sofronis P 31,43 Tromans, D. 31
Speck, J.S. 109 Turner, T. 224
Srolovitz, D.J. 89,108 Tvergaard, V. 2,3,9, 18, 367
AUTHOR INDEX 413

Ullner, C. 332 Wu, P. D. 17


Wu, C.H. 112
Vahdati, M. 337
Vanderbilt, D. 109
Vardoulakis, I. 171 Xia, L. 225,233
Vardoulakis, I.G. 169 Xia, Z. C. 17
Vehoff, H. 31,44,324 Xin, X.J. 87
Venkateswara Rao, K.T. 367,370 Xu, G. 90
Vieira, S 82 Xu, K. 362
Viggiani, G. 169 Xu, X.-P. 278,299
Voorhees, P.W. 109
Yamamoto, H. 2
Walker, B.E. 358 Yang, W. 358
Walston W.S. 31,44 Yau, J.F. 184,188
Wang, P. 362,367,370,372,378 Yoffe, E. 277
Wang J.-S 3l-32,35,39,41-42 Yokogawa, K. 31
Wang, S.S. 184,188 Yoshikawa, A. 31
Wang, Y. 114,396 Yu, J. 41,311-312
Weiland, H. 152-153
White, G.S. 332 Zdunek, A. B. 71
Wiederhorn, S.M. 332,346 Zeppenfield, P. 108
Wieserman, L.F. 152 Zhang T.-Y. 43,241
Wight, J.K. 402 Zhang, W. 184
Wilkins, B.J.S. 338 Zhou, S.J. 331
Williams, M.L. 237 Zhu, W. 161
Willis, J.R. 112 Zhu, J.Z. 337
Wilsdorf, H. 152 Zickgraf, B. 358
Wittig, F 50 Zienkiewicz, O.C. 337
Wong, T.F. 161,170 Zipse, H. 50,52-53,56,58
Wu, H.C. 391,394-395-396 Zureick, A.H. 123
This page intentionally left blank.
Subject Index

ABAQUS 7,50,58,65 Chemical potential 35


Acoustic Cleavage Griffith 42-43
waves 277 Cohesive
barrier 306 force law 275
Adhesion 71,72,74-76, 8l-83 crack model 387
map 81-83 models, 279-284,287
Alloy FeSi 31 Compact tension specimen 349
Amplification factor 147 Compaction, 160-161, 168-170
Analysis Composite
upper bond 2,24 ceramic matrix 361,375
failure localization 10 metal matrix 378
thermodynamic 32 Constitutive laws 145,151
finite element, 3D 1,3,6,13-14, 43,50, Contact 71-85
190, 295,351-352 sticking 80-81
Hall-Petch 88 frictionless 77
J-integral 290 scaling 72-76
asymptotic 252,271,273 maps 71,78-83
Anisotropic parameter R 3 elastic-plastic 71-85
Atomic Continuum damage model 277
diffusion 110 Coplanar slip 312,317,319,324,327
force microscopy 152,156 Crack
mobility 37 branching 277,296,298
Avalanche point 150 propagation 276,278
Bands 152,154 Crack growth 225-27,232-33,235
Bridged crack model 389 plane strain 259
Burgers vector 40,90,239 tensile 261
Cauchy resistance 234-35
principal value 97 anti-plane shear 244,250
symmetry 285,295 Crack tip 31,349
Cauchy-Born rule 279 plasticity 311
Cell Crack propagation 225-27,230,243-44
walls 147 Crack dynamics 275
physics 151 Crack opening 352
Cellular structure 147 Crack bifurcation, 298
Ceramic Crack growth rate 39
fiber 361 Crack opening displacement 381
particulate-reinforced 376 Crack initiation 33,35
416 SUBJECT INDEX

Crack opening displacement 387 Embedded


Crack fictitious 386 Atom method 40-41,88,93
Crack cohesive 386 process zone model 390,398
Crack bridged 386 Embrittlement,
Critical system self organizing 151 interfacial 32
Crystal orientation 313 Hydrogen 31
Damage tolerance 399,401 Energetics 110
Damage evolution 14,398 Energy
Deformation stacking fault 87
anti-plane shear 243 surface 112
plastic 25,43,381 elastic 110-111
Dielectric permittivity 354 density 229-230
Diffusion, 168, 167, 174. release rate 37,349,354
bulk 53,61 stacking 40
grain boundary 53,61 stacking fault 96,100-102,239
surface 53,61 unstable stacking 239
hydrogen 44 Epi1ayer108,110,112,116
Diffusivity 114 Etch pit 311,325,328
Dilatant, 161, 166-168, 171 factor character 150
Dilatational wave 295,297 Failure
Dislocation prediction 2,13-14
nucleation 37,237 criterion 291
emision 41,96, 237 indicator 291
Volterra 89,99 Fiber
screw 87,89 pull-out 381,396
Displacement field crack tip 319,321 reinforcement 365
Dissipation plastic work 23 bridging 386
Domain reorietation 358 Fiber-reinforced MMC 370
Double cantilever beam 226-28 Field intensity factor 358
Ductile particles 368 Finite element formulation
Effect shielding 42 Lagrangian 294
Elastic Finite strain 275,286
energy 110 Forming limit 2
space 123, 143 Fracture 349
field 123-143 process zone 31,223,226,228,
constant 354 236-37
Elastic-plastic, 171, 175 interface 244
Elasto-plastic 223-33,235,237 specific work of 224,226,237
Electrical enthalpy 350 mechanics 43,385
Ellipsoid 124
Ellipticity 292,294
SUBJECT INDEX 417

Fracture Kinetics segregation 38,44


extra work 227,229-32,236-37 L-integral 205,212,214
energy 398 Lames coefficients 126
resistance 375 Langmuir-McLean model 37
essential work of 223,225- Latent hardening 311,327-28
27,230-33,235-37 Lattice rotation 154, 321-22
Frankel model 239 Line-spring concept 387
Function Load-displacement 230,232,234-36
bond density 285 Load-displacement curve 314-15, 392
relaxation 155 Loading mode I 363-64
yield 20,27 Loading, axisymmetric 3
Gauss-Chebyshev quadrature 367 Local J scalar 208-210
Gibbs-Duhen relation 36 Local J vector 205-206,209-210,215
Global J scalar 208,217-218 Localization, 160-162,171, 173, 175-177
Global J vector 208,215 Localization flow 3
Grain coarsening 57 Localized ordering 152
Grain growth 54 Lomer-Cottrell locks 148,324
Graphie/epoxy 189 M-integral 183,188,205,212-213
Green body 49 Mass transport 110
Hilert distribution 60 Material
Hillert Law 57,59 piezoelectric 349,351,354
Hookes law 284-85 composite 385
Hysteretic loop 402 design 385,402
Inclusion rigid ellipsoidal 123 cementitious 385
Indentation 71-85 layered 244
maps 71,78-83 homogeneous 243-244,255,259,261
pre-stressed 80-81 elastic-plastic 243
master curves 77-78 tension softening 392
Inhomogeneity 123 characterization 385
Instability 165, 167, 174, 275,277-78, testing 398-403
300, 381 McLean isotherm 41
Interaction potential 290 Mechanism
Interface lock breaking 148
non-slipping 87 source 148
bimaterial 87 fracture resistance361
Cu/Ni 90 Mobility grain boundary 54
epitaxial 88 Model
crack 183 virtual internal bond 275, 279, 284,
Isotropic transversely 183 291- 292, 294,296
J-integral 205,350,353,385 388,393-395 constitutive 5l-58
J-integral modified 205,219 Dugdale 371
JR-curve 226,229,232-37 Moire 311
Keldysh-Sedove problem 366,373 Molecular dynamics 275
418 SUBJECT INDEX

Nanophases 108 Sheet forming 1


Newton method 353 Sheet metal rate-sensitive 3
Non-coplanar slip 312, 317, 319, 324, Sheet cell model 4,26
327 Similarity 71-76
Non-recoverable work 225 Single crystal 31,37
Nonlinear elastic 223-33, 235,237 copper 311
Notch sensitivity 399 iron 42
Numerical algorithms 294 Sintering aid 65
Partculate reinforcement 367 Slip 152-153
Pattern formation 146 active system 40
Peierls-Nabarro Model 238 bands 153-154
Peierls-Nabarro approach 40 elementary unit 152
Percolation 145 line 152
Periodic pattern 108 plane 40,101
Phase refining 109-110 trace 312
Phase coarsening 109-l10 Small scale
Plastic bridging 390
anisotropy 2 yielding 256,266
work accumulated 231 Soil, 160, 161, 163, 169, 177
flow 2 Solid porous plastc 2
zone 238 Solid Surface 108
Plasticity rigid-perfect 7 Solute segregation 34
PMMA 277,295 Solution upper bond 4
Polarization 358 Spanning cluster 149
Pore pressure 160-163, 165, 167, Sphere rigid 140
172-174, 177-180 Spheroid oblate 132
Pore fluid 160, 162, 166, 168, 172 Spheroid prolate 136
Potential function 124 Stiffness,
Powder technology 49 relation 143
Powder silicon carbide 58 tangent 277
Process simulation 50 Strain
Process zone 387,389 anti-plane 248
Rayleigh wave 299 cluster 147
Recoverable work 225,228,237 energy density 283,291
Rice-Thomson model 40 Green 292
Rock, 160-161, 163, 167-168, 169 Green Lagrangian 284
Rotation rigid particle 123- 142 hardening 3
Scaling law 149 Strain rate
Self-organization 109 macrscopic 21,25
Shear in-plane 7 plastic 20-21
Shear wave 295,300,302-304 shear 21
Sheet aluminum 17-18,21,29 Strength yield 19, 243
Sheet metal failure 3
SUBJECT INDEX 419

Stress Tensor elastic stiffness 284


Cauchy 286 Theory
critical shear 87 dilation plasticity 17,19
effective 27-28 Gurson type 18,21
eqibiaxial 287,289 density functional 93
intensity factor 354 continuum plasticity 18
mean 27 Thermodynamics segregation 44
Peierls 40 Threshold percolation 148,150-151
Piola-Kirchhoff 28l-82,288,291-92 Time localization 153
potentials 363 Transition ductile/brittle 37
plane 727 Translation rigid particle 123- 142
resolved shear 327 Transport 146
sintering 51,56 Universality class 148
surface 113 Unzipping event 148
yield 20 Viscosity shear 54
yield uniaxial 27 Viscosity bulk 55
Stress-displacement relation 92 Void coalescence 9,14,18
Stress/strain relations 145 Void growth 9,14,18
Stress-separation curve 32 Void nucleation 2,18
Stress-strain law, universal 153 Williams expansion 237
Stress-strain law 152 X-ray 311,321
Stress-strain curve 278 Yield criterion
Stretching, eqibiaxial 285 anisotropic 1,3,27,29
Stroh formulation 183, 194 Gursons 2-3,7-8,13-14
Stylus profilometer 311-12 Hills 1-3,11,13,
Surface stress 113-l14 Yield surface 2
Surface profile 320 curvature effect 10
Surface energy 111-112 Zone
T-stress 237 mist 277-78,289,298
deformation-softened 300
Temperature, Dugdale cohesive 378
ductile vs.brittle transition 39,41 bridge 362,364,366
Tension plastic 326
hydrostatc 7 hackle 277-78,289,298
biaxial 5,7 plastic 367
uniaxial 7,25-26 process 361

Вам также может понравиться