Вы находитесь на странице: 1из 12

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/220016362

High Magneto-Crystalline Anisotropic Core-Shell


Structured Mn0.5Zn0.5Fe2O4/PANI NanoComposites Prepared by in-situ Emulsion
Polymerization
Article in The Journal of Physical Chemistry C January 2012
Impact Factor: 4.77

READS

112

6 authors, including:
M. Abdullah Dar

VIVEK KUMAR VERMA

National Institute of Technology, Srinagar

Hindu College,University of Delhi

17 PUBLICATIONS 200 CITATIONS

47 PUBLICATIONS 388 CITATIONS

SEE PROFILE

SEE PROFILE

Jyoti Shah

W.A. Siddiqui

National Physical Laboratory - India

Jamia Millia Islamia

74 PUBLICATIONS 451 CITATIONS

81 PUBLICATIONS 1,385 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: VIVEK KUMAR VERMA


Retrieved on: 13 June 2016

Article
pubs.acs.org/JPCC

High Magneto-Crystalline Anisotropic CoreShell Structured


Mn0.5Zn0.5Fe2O4/Polyaniline Nanocomposites Prepared by in Situ
Emulsion Polymerization
M. Abdullah Dar,, R. K. Kotnala,*, Vivek Verma, Jyoti Shah, W. A. Siddiqui, and Masood Alam

National Physical Laboratory, Dr. K. S. Krishnan Road, New Delhi-110012, India


Department of Applied Science & Humanities, Jamia Millia Islamia University, New Delhi-110025, India

ABSTRACT: Nanosized Mn0.5Zn0.5Fe2O4 ferrite particles


were synthesized by a reverse microemulsion reaction process.
Coreshell nanocomposites composed of Mn0.5Zn0.5Fe2O4
nanocrystals and conjugated polymer polyaniline were successfully synthesized from a simple and inexpensive process, and in
situ chemical oxidative polymerization of aniline occurs.
Mn0.5Zn0.5Fe2O4 nanocrystals are in the micellar solution of
dodecyl benzene sulfonic acid (DBSA). It acts as both a surfactant and a dopant. These nanocomposites were subsequently
investigated for structural, morphological magnetic, and electromagnetic properties. Transmission electron microscopy (TEM) studies reveal the formation of coreshell type morphology
with ferrite nanocrystals (911 nm) as the center, while the polymer (PANI) formulates the outer shell of the composite.
A decrease in the blocking temperature of Mn0.5Zn0.5Fe2O4 nanocrystals after coating with PANI has been recorded, attributing
that the polymer matrix allows each ferrite nanoparticle to behave independently, indicating a decrease in the dipoledipole
interactions that will decrease the anisotropy energy barrier. Magnetic properties of the synthesized nanoferrite exhibit a very
high magneto-crystalline anisotropy energy value of 31.3 105 erg/cm3 at a blocking temperature of 90 K. The complex
permittivity, permeability, and microwave absorption properties of the composites were explored in the 8.212.2 GHz (X-band)
frequency range. The nanocomposites are observed to show high shielding effectiveness (SEA = 31.2 dB) as compared to pure
PANI, which strongly depends on dielectric loss, magnetic permeability, and volume fraction of Mn0.5Zn0.5Fe2O4 nanocrystals.
The high value of SEA suggests that these composites can be used as a promising electromagnetic absorbing material.

1. INTRODUCTION
With the advancement of electric and electronic industry, the
use of electronic products and telecommunication equipment
has increased due to which the problem of electromagnetic
interference (EMI) has attracted special consideration, as it
reduces the lifetime and competence of the instruments and
also affects the safety operation of many electronic devices. To
avoid these problems, all electronic equipment must be fortified
against electromagnetic destruction. Recently, extensive research has been done for the development of new microwaveshielding materials that have high efficiency, are lightweight,
and have high durability. This includes composites based on
polymers, like ferrite/polymer composites, metal/polymer
composites, epoxy-PANI/ferrite composites, and single-wall
carbon nanotube (SWCNT)/epoxy composites.13 Therefore,
demands to develop thinner electromagnetic (EM) wave
absorbers with wider bandwidths are ever increasing.4,5 EM
wave energy can be completely absorbed and dissipated into heat
through magnetic losses and dielectric losses if the characteristic
impedance of free space is matched with the input characteristic
impedance of an absorber. The microwave absorber can be used
to minimize the electromagnetic reflection from metal plates such
as aircraft, ships, tanks, walls of anechoic chambers, and electronic equipment.6,7 For the purpose of preparing a low-reflecting
2012 American Chemical Society

absorber in the desired frequency range, two fundamental conditions must be satisfied: the first is the incident wave that can
enter the absorber by the greatest extent, and the second is the
electromagnetic wave entering into the materials that can be
almost entirely attenuated and absorbed within the finite thickness of the absorber. Due to the high conductivity, environmental stability, and rather simple synthesis, polyaniline
(PANI) became the focus of attention for preparation of new
materials for the fabrication of industrial devices. In addition,
the PANI not only reflects but also absorbs electromagnetic
wave and may attain high levels of shielding performance.
Among the materials used in microwave absorbing applications,
ferrites exhibit an interesting behavior. Ferrites have been used
as absorbing materials in various forms, e.g., sheets, paints,
films, ceramic tiles, and powders loaded in matrix composites or
mixed with conducting materials. Hybridized systems of ferrites
and polymers have been the subject of considerable interest in
this research area. Such systems allow one to alter the electromagnetic properties of materials and thereby to meet requirements such as the matching condition and high absorption,
Received: June 16, 2011
Revised: January 12, 2012
Published: January 27, 2012
5277

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

giving a murky emulsion, followed by the addition of 5.55 wt % of


aqueous solution containing the precursor salts and 10.55 wt %
of isoamylalcohol under magnetic stirring, turning murky emulsion to transparent solution. The stirring was continued for
1 h, resulting in a stable reverse microemulsion (ME1). The
second reverse microemulsion solution (ME2) was prepared
with 0.1 M aqueous solution of NaOH as the water phase
under similar conditions. Both the reverse microemulsions were
then heated to 80 C, and the reverse microemulsion ME1 was
added dropwise to the reverse microemulsion ME2 under
constant magnetic stirring. The appearance of a blackish brown
color after a few minutes marks the completion of the reaction
and formation of the desired ferrite. The reaction mixture was
further stirred for 2 h on a magnetic stirrer at constant
temperature (80 C) and pH (9). The nanocrystals so obtained
were then collected by centrifugation (12 000 rpm/min). To
ensure complete removal of the surfactant, the powder was
subjected to several cycles of washing by methanol and doubledistilled water followed by centrifugation and finally dried in an
oven at 100 C for 36 h. A schematic representation for the
synthesis of nanocrystalline Mn0.5Zn0.5Fe2O4 is presented in
Figure 1.
2.2. In Situ Emulsion Polymerization of Aniline with
Mn0.5Zn0.5Fe2O4 Ferrite. Chemical oxidative emulsion
polymerization of aniline was carried out in the presence of
Mn0.5Zn0.5Fe2O4 nanocrystals in aqueous medium. The nanocrystals of Mn0.5Zn0.5Fe2O4 obtained above are homogenized in
0.3 M aqueous solution of dodecyl benzene sulfonic acid
(DBSA) to form a whitish brown emulsion solution. An appropriate amount of aniline (0.1 M) was added to the above solution and again homogenized for 2 h to form micelles of aniline
with Mn0.5Zn0.5Fe2O4 nanocrystals. The resulting micelles were
then polymerized at nearly 5 C for 12 h through oxidization
polymerization by ammonium persulfate ((NH4)2S2O8 (0.1 M))
indicated by the appearance of green color in the solution. The
product so obtained was demulsified by treating with an equal
amount of isopropyl alcohol. The precipitates were filtered out
and washed with methanol and then dried at 60 C. Several
composites having different weight ratios of aniline monomer to
ferrite (1:4, 2:3, 3:2, and 4:1) were synthesized, to study the
effect of the concentration of Mn0.5Zn0.5Fe2O4 nanocrystals on
the microwave absorption properties of the resulting PANI/
Mn0.5Zn0.5Fe2O4 nanocomposites. Besides this, pure polyaniline
doped with DBSA was also synthesized for their comparative
study. All these compositions having different weight ratios of
aniline monomer/ferrite are designated throughout this manuscript as a (Pure Ferrite), b (1:4), c (2:3), d (3:2), e (4:1), and
f (PANI).

thus reducing the reflection coefficient of electromagnetic waves


and broadening their operating frequency band. Ferrite
powders are often used as filling components for preparing
electromagnetic wave absorbers as the complex permeability of
such magnetic materials exhibits strong frequency dependence.8,9 Moreover, it has been found that insertion of ferrites
into a polymer matrix together with the amendment of the
shape, size, concentration, and the size distribution of ferrite
nanocrystals may strongly change the frequency dependence of
permeability.10,11 This change is mainly associated with
additional disorder in the fillermatrix interface. This disorder
is responsible for the magnetic discontinuity in the composites,
which leads to the broadening of the resonance bandwidth and
a shift of the resonance frequency to higher frequencies. However, at the same time, composite materials exhibit a
considerable decrease in permeability compared with bulk
ferrite materials. That is why one always seeks a compromise
between the value of permeability and the critical frequencies.
In the present study, we used Mn0.5Zn0.5Fe2O4 as a filling component for preparing electromagnetic wave absorber material
because of its fairly large magneto-crystalline anisotropy, high
initial permeability, and highest saturation magnetization among
all other spinel ferrites,12,13 which are important characteristics
for a material to act as an electromagnetic wave absorber.
Intensive efforts have been devoted to develop techniques for
the synthesis of nanocomposites; some of them are electrode
deposition,14 in situ polymerization,15 electrochemical synthesis,16 anodic oxidation,17 seeding polymerization,18 interfacial
polymerization,19 and electro-spinning,20 etc. However, limited
approaches have been used to develop a simple, mild, and efficient route to tune the properties of multicomponent ferrite/
PANI nanocomposites.21 Recently reverse microemulsion
polymerization has emerged as a useful strategy for the fabrication of polymerinorganic nanocomposites.22 The microemulsion technique allows control of both shape and size via
the structure of the surfactant assemblies. Most of the research
groups have studied Fe3O4/PANI nanocomposites;2325 however, very few reports have dealt with multicomponent ferrite/polyaniline nanocomposites.21,26 To the best
of our knowledge, there is no report focusing on the synthesis
of Mn0.5Zn0.5Fe2O4/PANI nanocomposites by a reverse microemulsion route.
In this work, we report the synthesis of highly monodisperse
Mn0.5Zn0.5Fe2O4 ferrite nanocrystals and nanocomposites of
conducting polymer polyaniline by incorporation of nanocrystals of Mn0.5Zn0.5Fe2O4 with superparamagnetic behavior by
in situ emulsion polymerization. The microstructures, morphology, and magnetic and electromagnetic properties of PANI/
Mn0.5Zn0.5Fe2O4 nanocomposites were investigated by means of
various experimental techniques.

3. CHARACTERIZATION
The crystallinity and phase analysis of the obtained samples
were investigated by an X-ray diffraction spectrophotometer
(XRD, Rigaku Miniflex, step size = 0.02) with Cu K radiation
of wavelength = 1.5406 . Fourier transform infrared (FT-IR)
spectra of the samples were recorded using a Perkin-Elmer
FTIR (Spectrum BX) in the region of 4002000 cm1. The
samples were prepared in the form of pellets with spectroscopic
grade KBr. The UVvisible spectra of PANI and its composites
were obtained with a Shimadzu 2250 UVvis spectrometer
using N-methyl-2-pyrrolidone (NMP) as solvent. The UV
wavelength ranged from 800 to 300 nm. The particle size
and morphology of the ferrite and composites were determined
using a high-resolution transmission electron microscope

2. EXPERIMENTAL SECTION
2.1. Synthesis of the Mn0.5Zn0.5Fe2O4 Ferrite Nanocrystals by the Reverse Microemulsion Method. To
synthesize Mn0.5Zn0.5Fe2O4 ferrite, an aqueous solution of
metal salts was prepared by mixing stoichiometric amounts of
0.1 M Fe(NO3)39H2O, 0.025 M Mn(NO3)26H2O, and 0.025 M
Zn(NO3)26H2O. A quaternary system of cyclohexane/CTAB/
isoamylalcohol/H2O was selected in this reverse microemulsion process. In this process, two solutions, namely,
(ME1) and (ME2), were prepared by a reverse microemulsion
technique. For the first reverse microemulsion solution (ME1),
CTAB (10.20 wt %) was added to cyclohexane (23.69 wt %),
5278

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

25 13 2 mm and inserted in the copper sample holder


connected between the waveguide flanges of the network
analyzer. Full two-port calibration is performed along with the
sample holder to neglect any loss and power redistribution due
to the sample holder.

4. RESULTS AND DISCUSSION


4.1. Formation Scheme. On the polymerization basis and
well supported by TEM measurements of the pure ferrite and
PANI/Mn0.5Zn0.5Fe2O4 (b) nanocomposite, the possible
mechanism for the formation of these coreshell nanocomposites has been proposed and portrayed schematically in
Figure 2. For the formation of this composite, emulsion
polymerization (oil in water type) has been carried out in which
the droplet of aniline (oil) emulsified with surfactant DBSA
containing Mn0.5Zn0.5Fe2O4 nanocrystals in a continuous phase
of water. A large amount of surfactant led to the formation of
micelle which in aqueous solution forms a nearly spherical or
globular aggregate with the hydrophilic head regions toward
the surrounding solvent, sequestering the hydrophobic tail
regions in the micelle center. Dodecyl benzene sulfonic acid has
been used as a stabilizer to prevent the chances of formation of
macroscopic crystals which act as a steric stabilizer and also an
effective doping agent.27 The presence of DBSA in the reaction
system increases the solubility of aniline due to enhanced
protonation of aniline and also improves the rate of the polymerization. Mn0.5Zn0.5Fe2O4 nanocrystals were first dispersed
into a solution of DBSA, and then aniline was added dropwise
and stirred to form a complex of anilineDBSA through the
interaction of aniline with the hydrophilic sulfonate group of
DBSA which separates the aggregated crystals into individual
crystals. Consequently, when ammonium persulfate solution
was added dropwise to the reaction system, the anilineDBSA
complex present on the surface of Mn0.5Zn0.5Fe2O4 nanocrystals polymerized, resulting in deposition of PANI on the
surface of Mn0.5Zn0.5Fe2O4 nanocrystals. Thus, coreshell
nanocomposites are engendered because of strong interactions
between the N atoms in the aniline polymer and those of
Mn0.5Zn0.5Fe2O4 nanocrystals through the electrostatic interactions. Moreover, it is important to note that the resulting
nanocomposites can form stable and high-quality coreshell
nanocrystals with uniform size as represented in Figure 2.
4.2. Structural and Morphology Analysis of PANI/
Mn0.5Zn0.5Fe2O4 Nanocomposites. X-ray diffraction (XRD)
was used for phase identification of the as-prepared Mn0.5Zn0.5Fe2O4 nanocrystals, PANI/Mn0.5Zn0.5Fe2O4 nanocomposite, and pristine PANI as shown in Figure 3. XRD peaks of
Mn0.5Zn0.5Fe2O4 nanoparticles confirm the single-phase spinel
structure with no other reflections. The diffraction peaks can be
well indexed to (220), (311), (400), (422), (511), and (440)
crystal plane of the cubic spinel. The lattice constant (ao) for
Mn0.5Zn0.5Fe2O4 nanocrystals was found to be 8.347 , which is
in agreement with the reported literature value. Figure 3 shows
the XRD patterns of the resultant polyaniline nanostructures.
Broad diffraction peaks centered at 2 = 20.1 and 25.7 can be
ascribed to the periodicity parallel and perpendicular to the
polymer chain. It indicated that the resulting polymer was in
the form of highly doped emeraldine salt and had good
crystallinity. As shown in Figure 3, it is clearly seen that both
the characteristic peaks of PANI centered at around 2 = 20.1
and 25.7 and Mn0.5Zn0.5Fe2O4 nanocrystals appear in the XRD
patterns of the nanocomposite. Also, the intensities of the
characteristic peaks of PANI became weaker after introducing

Figure 1. Schematic diagram for the synthesis of Mn0.5Zn0.5Fe2O4


ferrite nanocrystals by the reverse microemulsion technique.

(HRTEM) using the JEM 200CX model. The magnetic


measurements of pure ferrite and composites accomplished at
different temperatures were performed using a vibrating sample
magnetometer (Lakeshore, 7304). Electromagnetic shielding,
dielectric, and permeability measurements were carried out on
an Agilent E8362B Vector Network Analyzer in frequency
range of 812 GHz (X-band). Powder samples were compressed in the form of rectangular pellets having dimensions
5279

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

Figure 2. Schematic representation for the formation of Mn0.5Zn0.5Fe2O4/PANI coreshell nanocomposites.

Figure 3. X-ray diffraction pattern of pure PANI, 4:1 composite, and


pure Mn0.5Zn0.5Fe2O4 ferrite.

Mn0.5Zn0.5Fe2O4 nanocrystals into the polymer matrix, revealing


the decrease in the concentration of PANI in the nanocomposites.
The average diameter of the Mn0.5Zn0.5Fe2O4 nanocrystals calculated from the broadening of the XRD peak intensity using the
DebyeScherrer equation28 was found to be 6.2 nm. The
presence of a conducting shell encapsulating the magnetic and dielectric nanocrystals is helpful for the proper impedance matching,
which is necessary for enhancing the absorption of the electromagnetic wave.
The morphology and particle size distribution were determined by means of TEM. Figure 4(a) shows the representative
TEM image of the pure ferrite nanocrystals obtained from a
reverse microemulsion process, which produces a narrower and
spherical particle size distribution with average diameter of 6.8
nm. It was in close agreement with the XRD results, proving
that the nanocrystals were not significantly agglomerated. The
selected area electron diffraction (SAED) pattern as shown in
the inset of Figure 4(a) also reflected the crystallinity of the
ferrite particle formed with bright rings corresponding to
different planes (220), (311), (400), (422), (511), and (440) of
the spinel system and is totally in agreement with that of XRD.
Figure 4(b) represents the TEM image of the nanocomposite
having monomer/ferrite ratio of 1:4. From this figure it is

Figure 4. TEM images of (a) pure ferrite and (b) 1:4 monomer/ferrite
nanocomposite.

evident that the ferrite particles are embedded in the PANI


matrix to form the PANI/Mn0.5Zn0.5Fe2O4 composites with a
coreshell type structure. In comparison with the uncoated
ferrite particles, the PANI formed on the surface of ferrite
particles leads to the increase of the particle size as is evident
from Figure 4(b). The TEM image of PANI coated coreshell
material, Mn0.5Zn0.5Fe2O4PANI, illustrates that the dark core
material, and ferrite nanocrystals are coated by a light-colored
shell of PANI, due to the different electron penetrability.
5280

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

interaction takes place due to the capping of ferrite particles in


the polymer.
UVvisible spectra have been used to characterize the
interfacial interaction between the different samples of
polyaniline (PANI) and its composite with Mn0.5Zn0.5Fe2O4.
Figure 6 shows the UVvis spectra of these prepared PANI and

The crystallite size of particles in the nanocomposites was


found to be approximately 911 nm. The inset in Figure 4(b)
shows the SEAD patterns of the nanocomposite which are not
as perfect as that of ferrite, indicating the nanocomposite
consists of polycrystalline ferrite particles embedded in
amorphous PANI. These are in accordance with that of XRD
results. To establish coreshell structure of the nanocomposite,
a TEM image of higher magnification has been shown as the
inset in Figure 4(b).
The FTIR spectra of the Mn0.5Zn0.5Fe2O4 nanocrystals, PANI
DBSA, and composite of PANI/Mn0.5Zn0.5Fe2O4 nanocrystals
are shown in Figure 5. The pure ferrite sample (a) in Figure 5

Figure 6. UVvisible spectra of (b, c, d, e) composites of PANI/


Mn0.5Zn0.5Fe2O4 and (f) pure PANI.

PANI/Mn0.5Zn0.5Fe2O4 nanocomposites. The main peaks and


calculated energy bands are shown in Table 1. The conductive
Figure 5. FTIR spectra of (a) Mn0.5Zn0.5Fe2O4 ferrite (b, c, d, e)
composites of PANI/Mn0.5Zn0.5Fe2O4 ferrite and (f) pure PANI.

Table 1. Peak Positions and Calculated Energy Bands of


Pure PANI and PANI/Ferrite Nanocomposites
wavelength

exhibits the strong absorption bands that appeared at 462 and


584 cm1 (MO bonding) which are the characteristic bands
of Mn0.5Zn0.5Fe2O4 nanocrystals, enlightening the formation of
Mn0.5Zn0.5Fe2O4 nanocrystals in this experiment.29 The FTIR
spectra of Mn0.5Zn0.5Fe2O4 nanocrystals also exhibit a peak at
1635 cm1 attributed to the in-plane bending hydrogen-bonded
surface water molecules and hydroxyl groups.30,31 The PANI/
Mn0.5Zn0.5Fe2O4 composites have a similar FTIR spectrum, as
shown for samples (b), (c), (d), and (e) in Figure 5. There are
characteristic broad bands of Mn0.5Zn0.5Fe2O4 located at
around 462 and 584 cm1 in the IR spectrum, and PANI
occurs at 1570 and 1488 cm1 (CC stretching of the quinoid
and benzenoid rings), 1304 cm1 (CN stretching modes of
the benzenoid ring), 1246 cm1 (CH in-plane bending
modes), and 814 cm1 (CH out-of-plane bending modes).
The strong peak around 1134 cm1 is associated with vibration
modes of NQN (Q refers to the quinonic type rings) in
PANI/Mn0.5Zn0.5Fe2O4 ferrite composites.32 With an increase
in the PANI content, the intensity of the bands at 1570, 1488,
1304, and 1134 cm1 corresponding to PANI characteristics
increases distinctively, and the intensity of peaks at 462 and
584 cm1 corresponding to Mn0.5Zn0.5Fe2O4 was greatly diminished. These results designate that there is well wrapping of
Mn0.5Zn0.5Fe2O4 nanocrystals with PANI in the PANI/
Mn0.5Zn0.5Fe2O4 composites. Comparing the FTIR spectrum
of composite particles with that of the PANI, some peaks are
shifted to lower wave numbers, which is apparent from Figure 5,
confirming that some interaction of ferrite nanoparticles takes
place with polymer chains. There is no new band in FTIR
spectra of the nanocomposites, indicating that there is no
chemical reaction between ferrite and PANI; only a small

band gap (eV)

sample name

n*

n*

PANI
20% F
40% F
60% F
80% F

324
327
328
328
326

634
630
624
622
612

3.36
3.25
3.18
3.15
3.05

1.62
1.56
1.54
1.55
1.51

form of polyaniline doped with DBSA in N-methyl pyrrolidione


(NMP) shows two characteristic bands. The peak at about
324 nm is assigned to the * transitions within the benzenoid segment, and another peak near 634 nm is attributed
to the n* transitions within the quinoid structure.33 However, in the case of the polyaniline composite, two changes were
observed. First, a hypsochromic shift was observed for the band
from 634 to 612 nm, which was ascribed to the n*
transition. The reason behind this shifting may be the possible
interaction of the Mn0.5Zn0.5Fe2O4 with the polyaniline ring,
leading to the formation of a ferromagnetic composite. Second,
when the Mn0.5Zn0.5Fe2O4 content increases in different samples,
absorption spectra show the bathochromic shift for the band 324
326 nm. Generally, the absorption edge of organic semiconductors
can be determined by the following relation
h = (h Eg )n

(1)

where is the absorption coefficient; Eg is the band gap; and n is a


value that depends on the nature of the transition (1/2 for a direct
allowed transition or 2 for an indirect allowed transition). In this
case, n is equal to 1/2 for this direct allowed transition. The band
gap can be estimated from a plot of (h)2 versus photon (h)
energy as shown in the inset of Figure 6. The band gaps calculated
5281

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

by using the above eq 1 vary from 3.36 to 3.05 eV and 1.51 to


1.62 eV for the * transitions and n* transition of the
benzenoid ring, respectively. The decrease of band gap for the
PANI/Mn0.5Zn0.5Fe2O4 nanocomposite reveals that the electronic structure of PANI is affected by Mn0.5Zn0.5Fe2O4
nanocrystals.34 These results indicate that there could be interaction of Mn0.5Zn0.5Fe2O4 nanocrystals with amine and imine units
of PANI, causing changes in the electronic band.
4.3. Magnetic Properties. The room temperature hysteresis loop of all the samples was measured using a vibrating
sample magnetometer (VSM). Figure 7 shows the hysteresis

According to the StonerWohlfarth theory, magnetic anisotropy (EA) energy for single-domain crystals can be expressed as
EA = KV sin 2

(2)

where K is the magnetic anisotropy constant; V is the volume of


the nanoparticle; and is the angle between magnetization
direction and the easy axis of the nanoparticle. The anisotropy
represents the energy barrier to prevent the change of the
magnetization direction when the nanocrystal size is reduced to a
threshold value EA less than the thermal activation energy.
Consequently, the magnetization direction of the nanoparticle is
easily moved away from the easy axis due to thermal activation,
resulting in superparamagnetic behavior.
Figures 8 and 9 show the zero field cooling and field cooling
(ZFC-FC) curves recorded under an applied field of 50 Oe, for

Figure 7. MH loops of (a) Mn0.5Zn0.5Fe2O4 ferrite and (b, c, d, e)


composites of PANI/Mn0.5Zn0.5Fe2O4.

loops of Mn0.5Zn0.5Fe2O4 nanocrystals and PANI/Mn0.5Zn0.5Fe2O4


nanocomposites. It has been found that the hysteresis loop of
Mn0.5Zn0.5Fe2O4 nanocrystals could not be saturated even at
1 Tesla magnetic field. The saturation magnetization value of
4.3 emu/g for pure Mn0.5Zn0.5Fe2O4 nanocrystals (7 nm) at
room temperature is less than reported in the literature for
bigger particle size (25 nm).12,35 The hysteresis loops recorded
at room temperature for Mn0.5Zn0.5Fe2O4 nanocrystals exhibit
immeasurable remanence and coercivity, which would be
expected from these smaller size equiaxial-shaped nanocrystals.
This proves that the nanocrystals obtained from reverse microemulsion were superparamagnetic at room temperature.36,37
The magnetic properties in the composites were inherited from
the magnetic Mn0.5Zn0.5Fe2O4 nanocrystals.
The hysteresis loop for the nanocomposite gets saturated
with 1 Tesla on increasing the PANI content, indicating their
prominent ferromagnetic behavior. The interactions between
the polymer and ferrite nanocrystals play a significant role in
attaining the magnetic nature of these nanocomposites. The
interaction between PANI and Mn0.5Zn0.5Fe2O4 nanocrystal
increases due to H-bonding between the O of ferrite and
hydrogen of N of PANI.38,39 Hydrogen bonding between N
and O restricts the free magnetic spin orientation available on
the surface of the ferrite particle, and it may be responsible for
turning superparamagnetic behavior (nonsaturation of magnetic moment) into ferromagnetic behavior (saturation of magnetic moment) of the nanocomposite. Also, increased interaction between the polymer and ferrite nanocrystal decreases the
particleparticle exchange interaction and allows easy alignment of nanocomposite particles with applied magnetic field,
resulting in ferromagnetic nature.40

Figure 8. ZFC-FC response of sample a in the temperature range of


80310 K.

Figure 9. ZFC-FC response of sample b in the temperature range of


80310 K.

sample a and b synthesized via the reverse microemulsion


method. From the curves, it is clearly observed that bifurcation
of the ZFC-FC curves at a certain temperature TSEP is one of
the characteristic features of the superparamagnetic system.
However, at the same time a broad maximum is observed in the
ZFC curves at a slightly lower temperature than TSEP called
blocking temperature (TB). Under ZFC conditions, the nanocrystals
become magnetically frozen, and their magnetic moment could not
flip and align rapidly along the direction of the applied field. The
spins are able to flip only above the blocking temperature (TB),41
and the ZFC magnetization then coincides with the FC
5282

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

magnetization. At temperatures above TB, the thermal energy,


characterized by KBT, is larger than the magnetic anisotropy
energy barrier, and thus the material behaves superparamagnetic following the CurieWeiss law. The blocking temperature
of sample a is found to be 90 K, and for sample b blocking
temperature is expected to be <80 K as apparent in Figure 8
and Figure 9. A decrease in the blocking temperature after
coating with PANI can show that the polymer matrix allows
each ferrite nanoparticle to behave independently, indicating a
decrease in the dipoledipole interactions that will decrease the
anisotropy energy barrier. This feature is reliable with the other
experimental results as reported in the literature.42 The
behavior of superparamagnetic nanocrystals can be described
by the NeelBrown equation as

= o exp(EA /kBT )

Figure 11. MH curves of sample b at different temperatures (80


310 K).

(3)

where is the superparamagnetic relaxation time; o is a


relaxation time constant (1010 s); and T is the temperature
for a measurement time of, say, 60 s.
EA = ln(/o)kBT = 25kBT

saturation magnetization is observed to increase with respect to


temperature from 310 to 80 K. The nanocrystals indicated
remanent magnetization due to the hysteretic features that have
been exhibited for both the samples as shown in Figures 10 and 11.
The nanocrystals do not have adequate thermal energy to
attain complete thermal equilibrium with the applied field
during the measurement time, hence hysteresis appears. At
temperatures higher than blocking temperature, the hysteresis
disappears. The small magnetization value of 4.3 emu/g for
sample 'a' can be explained on the basis of cation distribution in
these ferrite nanocrystals. In nanoscale ferrites, cations occupy
lattice sites by a certain degree against their preferences in bulk
materials, and this degree (of inversion) is dependent on
particle size. As the crystal size decreases, there is a change in
the degree of inversion parameters; i.e., there are more Mn2+
ions occupying A-sites and also more Zn2+ ions occupying
B-sites, resulting in lower magnetization values for these
nanocrystals. The lower values of magnetization compared to
the multidomain bulk ferrite can also be attributed to surface
effects due to the finite size of nanocrystallites, which in turn
leads to a noncolinearity of magnetic moments on their surfaces. It can be explained in terms of the coreshell morphology of the nanocrystals which consists of ferri-magnetically
aligned core spins and a spin-glass like surface layer.44,45 If this
surface layer is absent, the magnetization of the nanocrystals
would saturate with an increase in applied field. Up to a particular magnetic field, the core magnetic moments align with
the applied magnetic field. At some stage, the resource of the
core mode of the magnetization response is exhausted, and
the core magnetization' of the system is saturated in a usual
Langavin-like way. Beyond this stage, any increase in the magnetic field on the nanocrystals has an effect only on the surface
layer of the nanocrystals, thus the increase in the magnetization
on them slows down. This specific state of the coreshell
morphology entails a virtual absence of magnetic saturation and
a very low coercivity value. The origin of surface spin disorder
for ferrite nanocrystals may be due to broken exchange bonds,
high anisotropy layer on the surface, or a loss of the long-range
order in the surface layer. These effects are particularly strong
in the case of ferrites because of the superexchange interactions through the oxygen ions. The small magnetization values
4.3 emu/g are difficult to justify considering surface effects
only.46,47 The existence of some degree of spin canting in the
whole volume of the particle, in addition to the cation distribution and disordered surface layer, could be an alternative

(4)

For a particle with uniaxial anisotropy EA = KV, the condition


for superparamagnetism becomes KV = 25kBT; K is the
magneto-crystalline anisotropy. Therefore, the blocking temperature should approximately satisfy the relation

TB = KV /25kB

(5)

The anisotropy is larger for nanocrystals than the bulk value,


and it increases with the decrease in particle size. Assuming that
the particles are spherical, the magneto-crystalline anisotropy
constant K is obtained as 31.3 105 erg/cm3 for TB = 90 K,
which is quite larger than the bulk ferrites. The nanocrystals
display the superparamagnetic relaxation when the energy
barriers of magnetic anisotropy are overcome.
The ZFC-FC curve separation for all the ferrites can be
described as high-field irreversibility at lower temperature and is
ascribed to the freezing of the disordered surface spins. The
high reversibility associated with the blocking process of the
magnetic nanocrystals is expected to disappear for applied fields
of a few KOe when the anisotropy fields of the nanocrystals are
surpassed and nanocrystals reach a state of saturation. However,
it is suggested that high-field reversibility is a strong function of
particle size for nanocrystals.43
The variation of magnetic behavior with applied field at
different temperatures (80310 K) for samples a and b is
shown in Figure 10 and Figure 11, respectively. The value of

Figure 10. MH curves of sample a at different temperatures (80310 K).


5283

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

explanation for this additional decrease of the magnetization. It


also seems that, in addition to the surface effect, the order
disorder characteristic of the samples also has a strong influence
on the final value of the saturation magnetization. These results
would explain the low magnetization values reported for very
small crystals. For both the samples (a and b), complete
saturation is not achieved even up to a field of 10 kOe. This
behavior is closely related with the existence of a magnetically
disordered surface layer in which direct competition of
exchange interactions between surface spins takes place.47 An
increase in the magnetization values of sample a from 4.3 to 23
emu/g at 310 and 80 K, respectively, has been observed due to
the decrease in thermal energy. Therefore, magnetic moments
will align along the external magnetic field direction leading to
an increase in the net magnetization. The low values of coercivity for Mn0.5Zn0.5Fe2O4 than nanocomposites (Figure 12)

Figure 13. MT curve for pure Mn0.5Zn0.5Fe2O4 nanoferrite.

4.4. Microwave Absorption Properties. The EMI


shielding effectiveness (SE) of a material can be expressed as
SE (dB) = 10 log(Pt/Po)

(6)

where Pt and Po are the transmitted and incident electromagnetic power, respectively. For a shielding material, total
SE = SER + SEA + SEM, where SER and SEA are the shielding
effectiveness due to reflection and absorption, respectively.
SEM is multiple reflection effectiveness inside the material,
which can be negligible when SE >10 dB. To express the
reflectance and effective absorbance we use the form of 10
log(1 R) and 10 log(1 Aeff) in decibel (dB), respectively,
hence the terms SER and SEA can be described as SER = 10
log(1R) and SEA = 10 log[T/(1 R)].
Figure 14 shows the response of EMI shielding effectiveness
(SE) of different nanocomposites with frequency having differ-

Figure 12. Variation of coercivity of pure Mn0.5Zn0.5Fe2O4 ferrite and


1:4 monomer/Mn0.5Zn0.5Fe2O4 nanocomposite (80%) with temperature.

are expected due to the particleparticle interactions among


the nanocrystals owing to their extremely small size. The
proximity of nanocrystals has a large affect on the hysteresis as
they either become increasingly exchange coupled or show
magnetostatic interactions with decreasing distance between
the nanocrystals.48 Thus, the magnetization shows an increase
on lowering the temperature, but the coercivity (Figure 12) in
this case follows an opposite trend; i.e., it decreases with a
decrease in measuring temperature. Nevertheless, the decrease
in coercivity is expected due to the lower anisotropy, indicating
the highly superparamagnetic nature of the nanocrystals
obtained from reverse microemulsion.49 The Curie temperature
investigated for the ferrite samples has been obtained from the
MT curve using a vibrating sample magnetometer as shown
in Figure 13. The value of TC is found to be 190 K which is
much less as compared to the multidomain nanocrystals. This
behavior can be ascribed to the variation in inversion degree of
cations. In nanoscale ferrite, cations occupy lattice sites by a
certain degree against their preferences in bulk materials, and
this degree of inversion is dependent on particle size. The Zn
incorporation reduces the number of Fe3+ ions on A-sites, which
in turn weakens the superexchange interaction between the A
and B sublattices. Therefore, for Mn0.5Zn0.5Fe2O4 nanocrystals
prepared by the reverse microemulsion process, there are more
Mn2+ ions occupying A-sites and also more Zn2+ ions occupying octahedral B-sites, resulting in lower TC values compared to
bulk ferrite.

Figure 14. Shielding effectiveness of (b, c, d, e) composites of PANI/


Mn0.5Zn0.5Fe2O4 and (f) pure PANI in the X-band frequency range.

ent weight ratios of the aniline monomer/Mn0.5Zn0.5Fe2O4. As


compared to the lower SE of PANIDBSA, the SE of the
composite first increases with the ferrite content exhibiting
excellent frequency stability in the measured frequency range
and then further decreases. The contribution to the SE values
mainly comes from the absorption rather than reflection in the
case of PANI/Mn0.5Zn0.5Fe2O4 nanocomposites. It is clear from
Figure 14 that composite d has the higher SEA of 31.2 dB with
insignificant SER of 4.8 dB as compared to the SEA value
of 28 dB and reflection SER of 6 dB for the PANIDBSA.
5284

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

Figure 15. Real (a) and imaginary (b) permittivity response of all samples in the X-band frequency range.

Figure 16. Real (a) and imaginary (b) permeability response of all composite samples in the X-band frequency range.

reveals that the material possessing higher conductivity and


magnetic permeability can achieve better absorption properties.
The scattering parameters S11 and S21 have been used for
determining the complex permittivity and the permeability of
PANI/Mn0.5Zn0.5Fe2O4 composites using the Nicholson, Ross,
and Weir method.32 The real part of permittivity is mainly
associated with the amount of polarization occurring in the
material, and the imaginary part is associated with the dissipation of energy. Both the values of the real and imaginary part
of permittivity are found to decrease with respect to frequency
for all samples as shown in Figure 15(a) and (b), respectively.
In conjugated polymers, strong polarization occurs due to the
presence of the polaron/bipolaron system that is mobile and
free to move along the chain, and others are bound charges
(dipoles) which have only restricted mobility and account for
strong polarization in the system.51 When the frequency of the
applied field is increased, the dipoles present in the system
cannot reorient themselves fast enough to respond to applied
electric field, and as a result the dielectric constant decreases.52
While incorporating Mn0.5Zn0.5Fe2O4 in the polyaniline matrix,
a significant increase in the real and imaginary part of complex
permittivity was observed as compared to pristine PANI.
The higher values of and are owing to more interfacial
polarization due to the presence of insulating Mn0.5Zn0.5Fe2O4
nanocrystals, consequently leading to more shielding effectiveness due to microwave absorption.
Figure 16(a) and (b) shows the variation in the real and
imaginary parts of permeability with frequency for different composites having different weight ratios of monomer/
Mn0.5Zn0.5Fe2O4. The real part of the permeability remains

These values are higher than reported in the literature for


similar ferrite composites (10 dB)13 and some other PANIFe2O3 composites (8.8 dB).50 The increase in the absorption
part with the addition of ferrite nanocrystals may be attributed
to the higher dielectric and magnetic losses observed in the
polymer composites, while in the case of PANIDBSA, only
dielectric losses contributed to the SEA values.51 According to
classical electromagnetic theory the EMI shielding effectiveness
can be expressed as
SE (dB) 10 log(ac /16ro) + 20(d /)log e

(7)

The first term in eq 7 is shielding effectiveness due to


reflection, and the second is due to the absorption of the
electromagnetic wave. Therefore, from eq 7 shielding effectiveness due to reflection and absorption can be expressed by
eqs 8 and 9, respectively.
SER (dB) = 10 log(ac /16ro)

(8)

SEA (dB) = 20(d /)log e = 20d(rac /2)1/2 log e


(9)

where d is the thickness of the shield; r is the magnetic


permeability; is the skin depth; ac = o is the frequencydependent conductivity; is the imaginary part of permittivity
(loss factor tan ); is the angular frequency; and o is the
permittivity of the free space. From eqs 8 and 9, it is observed
that with the increase in frequency the SEA values increase
while the contribution to the reflection loss decreases.
Dependence of SEA and SER on conductivity and permeability
5285

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

Notes

constant with a little fluctuation, whereas the imaginary part


shows an increase in the measured frequency range. The permeability was observed to increase for a higher percentage of
ferrite in the polymer matrix. The nanocomposite e has the
highest permeability value of 0.82, having a 4:1 weight ratio of
monomer/Mn0.5Zn0.5Fe2O4. The higher value of has been
observed for composite d as compared to others, which confirms
the existence of greater magnetic losses in this nanocomposite.
When the frequency of the applied field increases, the magnetic dipole tries to rotate with the frequency, but at higher
frequency due to strong anisotropy, the induced magnetization
(B) lags behind the applied field (H) which results in magnetic
losses. The larger the anisotropy, the higher the difference in B
and H, which results in more magnetic losses in the material.53
In magnetic nanocrystals, the rotation of domains becomes difficult due to the effective anisotropy (magneto-crystalline anisotropy and shape anisotropy). The effective anisotropy coefficient
can be enhanced significantly with decreasing particle size due to
the surface effect and microstructure defects.54,55 Therefore, an
increase in Mn0.5Zn0.5Fe2O4 nanocrystal content in the polymer
matrix leads to higher magnetic losses, which in turn improves
absorption of the electromagnetic microwaves.

The authors declare no competing financial interest.

ACKNOWLEDGMENTS
The authors are grateful to Director, National Physical Laboratory, New Delhi, for providing constant encouragement and
motivation to carry out this work. The authors are grateful to
Dr. S. K. Dhawan, NPL New Delhi, for providing the facility of
EM absorbing measurements. M. Abdullah Dar is also thankful
to University grants commission for their financial assistance
through grant No. F1-17.1/2011/MANF-MUS-JAM-638.

CONCLUSIONS
In this work, we have reported the synthesis of superparamagnetic Mn0.5Zn0.5Fe2O4 nanocrystals prepared by the reverse
microemulsion reaction method. Coreshell nanocomposites
composed of Mn0.5Zn0.5Fe2O4 ferrite nanocrystals as core and
conjugated polymer polyaniline as shell were successfully
synthesized by in situ emulsion polymerization of aniline with
Mn0.5Zn0.5Fe2O4 nanocrystals in dodecyl benzene sulfonic acid
(DBSA) which acts as both a surfactant and a dopant. Transmission electron microscopy (TEM) studies reveal the
formation of coreshell type morphology with ferrite nanocrystals (911 nm) as the center, while the polymer (PANI)
formulates the outer shell of the composite. UVvisible spectra
have been used to characterize the interfacial interaction between the different samples of polyaniline (PANI) and its composite with Mn0.5Zn0.5Fe2O4 nanocrystals. The magnetization
of the composite shows an increase on lowering the measuring
temperature, while the coercivity follows a reverse trend; i.e.,
it decreases with the decrease in measuring temperature. It
was observed that on nanocomposite formation with PANI
Mn0.5Zn0.5Fe2O4 nanocrystals undergo a transition from being
superparamagnetic to ferromagnetic. A decrease in the blocking
temperature of Mn0.5Zn0.5Fe2O4 after coating with PANI attributes
that the polymer matrix allows each ferrite nanoparticle to behave independently. The nanocomposites exhibited
strong microwave absorption properties in the X-band frequency range of 31.2 dB at 12 GHz with a minimal reflection
loss of 4.8 dB for sample (d) having a 3:2 ratio of monomer/
ferrite. The high absorption properties mainly resulted from the
high dielectric and magnetic losses in the nanocomposites and
depend on the concentration of Mn0.5Zn0.5Fe2O4 nanocrystals
in the polymer composite. As a result, nanocomposites of
PANI/ferrites with coreshell type morphology are promising
as new types of microwave absorptive materials that in a wide
range of frequencies maintain strong absorption.

REFERENCES

(1) Abbas, S. M.; Chatterjee, R.; Dixit, A. K.; Kumar, A. V. R.; Goel,
T. C. J. Appl. Phys. 2007, 101, 074105.
(2) Che, R.; Peng, L. M.; Duan, X.; Chen, Q.; Liang, X. Adv. Mater.
2004, 16, 401.
(3) Huang, Y.; Li, N.; Ma, Y.; Du, F.; Li, F.; He, X.; Lin, X.; Gao, H.;
Chen, Y. Carbon 2007, 46, 1614.
(4) Meshram, M. R.; Agrawal, N. K.; Sinha, B.; Misra, P. S. J. Magn.
Magn. Mater. 2004, 271, 207.
(5) Sugimoto, S.; Kondo, S.; Okayama, K.; Book, D.; Kagotani, T.;
Homma, M. IEEE Trans. Magn. 1999, 35, 3154.
(6) Feng, Y. B.; Qiu, T.; Shen, C. Y.; Li, X. Y. IEEE Trans. Magn.
2006, 42, 36.
(7) Abbas, S. M.; Chandra, M.; Verma, A.; Chatterjee, R.; Goel, T. C.
Composites, Part A 2006, 37, 2148.
(8) Nakamura, T.; Miyamoto, T.; Yamada, Y. J. Magn. Magn. Mater.
2003, 256, 340.
(9) Yusoff, A. N.; Abdullah, M. H. J. Magn. Magn. Mater. 2004, 269,
271.
(10) Tsutaoka, T. J. Appl. Phys. 2003, 93, 2789.
(11) Moucka, R.; Lopatin, A. V.; Kazantseva, N. E.; Vilcakova, J.;
Saha, P. J. Mater. Sci. 2007, 42, 9480.
(12) Gubbalaa, S.; Nathania, H.; Koizolb, K.; Misra, R. D. K. Phys. B
2004, 348, 317.
(13) Lee, S. P.; Chen, Y. J.; Ho, C. M.; Chang, C. P.; Hong, Y. S.
Mater. Sci. Eng., B 2007, 143, 1.
(14) Sazou, D. Synth. Met. 2001, 118, 133.
(15) Sunderland, K.; Brunetti, P.; Spinu, L.; Fang, J.; Wang, Z.; Lu,
W. Mater. Lett. 2004, 58, 3136.
(16) Jarjayes, O.; Fries, P. H.; Bidan, C. Synth. Met. 1995, 69, 343.
(17) Yan, F.; Xue, G.; Chen, J.; Lu, Y. Synth. Met. 2001, 123, 17.
(18) Thanpitcha, T.; Sirivat, A.; Jamieson, A. M.; Rujiravanit, R.
Synth. Met. 2008, 158, 695.
(19) Li, J.; Jia, Q.; Zhu, J.; Zheng, M. Polym. Int. 2008, 57, 337.
(20) Jun, J. H.; Cho, K.; Yun, J.; Suh, K. S.; Kim, T. Y.; Kim, S. Org.
Electron. 2008, 9, 445.
(21) Wu, K. H.; Shin, Y. M.; Yang, C. C.; Ho, W. D.; Hsu, J. S.
J. Polym. Sci., Part A: Polym. Chem. 2006, 44, 2657.
(22) Xie, G.; Zhang, Q.; Luo, Z.; Wu, M.; Li, T. J. Appl. Polym. Sci.
2003, 87, 1733.
(23) Xiao, Q.; Tan, X.; Ji, L.; Xue, J. Synth. Met. 2007, 157, 784.
(24) Bao, L.; Jiang, J. S. Phys. B: Phys. Condens. Matter 2005, 367,
182.
(25) Xuan, S.; Wang, Y. X. J.; Leung, K. C. F.; Shu, K. J. Phys. Chem.
C 2008, 112, 18804.
(26) Zhong, A. H.; Hong, X. Z.; Chao, K.; Yu, Y. Y.; Xiu, L. S.; Li,
J. R. J. Mater. Sci.: Mater. Electron. 2006, 17, 859.
(27) Han, D.; Chu, Y.; Yang, L.; Liu, Y.; Lv, Z. Colloids Surf., A 2005,
259, 179.
(28) Abdullah Dar, M.; Batoo, K. M.; Verma, V.; Siddiqui, W. A.;
Kotnala, R. K. J. Alloys Compd. 2010, 493, 553.
(29) Kotnala, R. K.; Abdullah Dar, M.; Verma, V.; Siddiqui, W. A.
J. Magn. Magn. Mater. 2010, 322, 3714.
(30) Dallas, P.; Stamopoulos, D.; Boukos, N.; Tzitzios, V.; Niarchos,
D.; Petridis, D. Polymer 2007, 48, 3162.

AUTHOR INFORMATION

Corresponding Author

*Tel.: +91-11-45608599. E-mail: rkkotnala@nplindia.org.


5286

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

The Journal of Physical Chemistry C

Article

(31) Garcia-Cerda, L. A.; Montemayor, S. M. J. Magn. Magn. Mater.


2005, 294, 43.
(32) Gairola, S. P.; Verma, V.; Kumar, L.; Abdullah Dar, M.;
Annapoorni, S.; Kotnala, R. K. Synth. Met. 2010, 160, 2315.
(33) Jiang, J.; Li, L.; Xu, F. J. Phys. Chem. Solids 2007, 68, 1656.
(34) Pal, M.; Hirota, K.; Sakata, H. Phys. Status Solidi A 2003, 196,
396.
(35) Makovec, D.; Kodre, A.; Arcon, I.; Drofenik, M. J. Nanopart. Res.
2009, 11, 1145.
(36) Upadhayay, T.; Upadhayay, R. V.; Mehta., R. V. Phys. Rev. B
1997, 55, 5585.
(37) Xaun, S.; Hao, L.; Jiang, W.; Gong, X.; Hu, Y.; Chen, Z. J. Magn.
Magn. Mater. 2007, 308, 210.
(38) Li, Y. X.; Zhang, H. W.; Liu, Y. L.; Wen, Q. Y.; Li, J.
Nanotechnology, 2008, 19.
(39) Li, L. C.; Jiang, J.; Xu, F. Mater. Lett. 2007, 61, 1091.
(40) Kumar, S.; Singh, V.; Aggarwal, S.; Mandal, U. K.; Kotnala, R. K.
Compos. Sci. Technol. 2010, 70, 249.
(41) Neel, L. Rev. Mod. Phys. 1953, 25, 293.
(42) Lyon, J. L.; Fleming, D. A.; Stone, M. B. Nano Lett. 2004, 4, 719.
(43) Zhang, Z.; Zhang, Y. D.; Xiao, T. D.; Ge, S.; Wu, M.; Hines,
W. A.; Budnick, J. I.; Gromek, J. M.; Yacaman, M. J.; Troiani, H. E.
Mater. Res. Symp. Proc. 2002, 3.5.1, 703.
(44) Hasmonay, E.; Depeyrot, J.; Sousa, M. H.; Tourinho, F. A.;
Bacri, J. C.; Perzynski, R.; Raikher, Y. L.; Rosenman, I. J. Appl. Phys.
2000, 88, 6628.
(45) Morr, A. H.; Haneda, K. J. Appl. Phys. 1981, 52, 2496.
(46) Vollath, D.; Szabo, D. V.; taylor, R. D.; Willis, J. O. J. Mater. Res.
1997, 12, 2175.
(47) Martinez, B.; Obradors, X.; Balcells, L. L.; Rouanet, A.; Monty,
C. Phys. Rev. Lett. 1998, 80, 181.
(48) Verdes, C.; Thompson, S. M.; Chantrell, R. W.; Stancu, A. Phys.
Rev. B 2002, 65, 174417.
(49) Roca, A. G.; Marco, J. F.; Morales, M. D. P.; Serna, C. J. J. Phys.
Chem. C 2007, 111, 18577.
(50) Singh, K.; Ohlan, A.; Bakhshia, A. K.; Dhawan, S. K. Mater.
Chem. Phys. 2010, 119, 201.
(51) Ohlan, A.; Singh, K.; Chandra, A.; Dhawan, S. K. Appl. Mater.
Interfaces 2010, 2, 927.
(52) Phang, S. W.; Hino, T.; Abdullah, M. H.; Karamoto, N. Mater.
Chem. Phys. 2007, 104, 327.
(53) Ishino, K.; Narumiya, Y. Ceram. Bull. 1987, 66, 1469.
(54) Dimitrov, D. A.; Wysin, G. M. Phys. Rev. B 1995, 51, 11947.
(55) Shilov, V. P.; Bacri, J. C.; Gazeau, F.; Gendron, F.; Perzynski, R.;
Raikher, Y. L. J. Appl. Phys. 1999, 85, 6642.

5287

dx.doi.org/10.1021/jp205652d | J. Phys. Chem. C 2012, 116, 52775287

Вам также может понравиться