You are on page 1of 8

Journal of Alloys and Compounds 566 (2013) 5461

Contents lists available at SciVerse ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Inuence of Co2+ distribution and spinorbit coupling on the resultant


magnetic properties of spinel cobalt ferrite nanocrystals
S.R. Naik a, A.V. Salker a,, S.M. Yusuf b, S.S. Meena b
a
b

Department of Chemistry, Goa University, Goa 403 206, India


Solid State Physics Division, Bhabha Atomic Research Centre, Mumbai 400 085, India

a r t i c l e

i n f o

Article history:
Received 11 January 2013
Received in revised form 15 February 2013
Accepted 24 February 2013
Available online 16 March 2013
Keywords:
Oxide materials
Sol gel process
Spin orbit effect
Magnetically ordered materials
Mossbauer
TEM

a b s t r a c t
Superparamagnetic properties of the cobalt ferrite nanocrystals have been demonstrated. The signicance of the solgel autocombustion method in preparation of cobalt ferrite oxide in the nano range
(3040 nm) has been very well complimented with the structural, dimensional and morphological techniques, such as X-ray diffraction technique, Transmission Electron Microscopy and Scanning Electron
Microscopy. The valence states of the metal ions and single phase formation of the polycrystalline oxide
have been conrmed with the help of X-ray photoelectron spectroscopy and Raman spectroscopy. The
distribution of the Fe3+ ions in the tetrahedral and octahedral lattice sites has been illustrated with the
help of the Mssbauer spectroscopy that shows ve sextets, indicating occupancies of one tetrahedral
and four octahedral sites by Fe3+ ions. Hyperne elds of 51.29, 48.74, 46.78, 43.58 and 48.59 Tesla,
respectively in CoFe2O4 have been found for four octahedral and one tetrahedral site respectively, at
ambient temperature. The magnetic measurements MH and MT demonstrate a change in the magnetic
moment and a superparamagneticferrimagnetic transition at 235 K in the ferrite system.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Magnetic spinel ferrites are of great technological interest and
have grabbed the attention of many researchers due to their fascinating electromagnetic properties. The ease of preparation and the
stability of these materials under various conditions have added lot
of signicance for their use in the industry. The novelty observed in
the physico-chemical and the electromagnetic properties due to
changes in the particle dimension and composition has encouraged
researchers around the globe to synthesize these materials [17].
The exotic magnetic, electronic and magneto-optical properties
of the cobalt ferrite (CoFe2O4) have prioritized them in the manufacturing of magnetic recording devices, magnetic refrigeration,
catalysis, sensors and magnetic uids [8,9]. Magnetic nanocrystals
have been of great interest over the past several years for the fundamental understanding of nanomagnetism and for their technological and biomedical applications [812]. Single phase
magnetic ferrites and their nanocomposites nd tremendous application as antimicrobial agents and vehicle for drug delivery, thus
proving their vitality even in the non-technological eld, thereby
expanding their horizon of application [1315]. The magnetic
properties of nanocrystals vary greatly with changing dimensions
and superparamagnetism is a typical example of such a size-

Corresponding author. Tel.: +91 832 6519315; fax: +91 832 2452889.
E-mail addresses: sal_arun@rediffmail.com, sav@unigoa.ac.in (A.V. Salker).
0925-8388/$ - see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jallcom.2013.02.163

dependent behavior at the nanometer scale. Different applications


such as technical and medical, demands different set of magnetic
properties. Therefore, approaches aiming at the synthesis of ferrite
oxides with uniform particle size are essential for controlling the
magnetic properties to satisfy the requirements for such applications. Systematic studies on the correlations between magnetic
properties and the chemical compositions of nanocrystals will generate invaluable insight towards the fundamental understanding of
magnetic properties and consequently enable to identify suitable
candidates of magnetic nanocrystals for various applications. Over
the years, the magnetic nanoparticles that have been studied are
mostly the iron containing oxides, particularly magnetite (Fe3O4)
which could be due to its easy availability. Now, with the advancement in nanoscience and nanotechnological research, new types of
magnetic nanocrystals have become available.
CoFe2O4 belongs to the crystal family of spinel ferrites (AB2O4).
The magnetic properties of these oxides are mainly dominated by
the distribution and the magnetic interaction of the cations in the
two lattices, i.e., tetrahedral (A) and the octahedral (B). Because of
the difference in the strengths of magnetic interactions at these
two lattices of the spinel oxides, the magnetic properties possessed
by CoFe2O4 nanocrystals vary with change in the cation distribution, thereby generating scope for the various applications. CoFe2O4 crystallizes as random spinel which can be represented as
2
3
3
(Co2
x Fe1x )Co1x Fe1x O4 , where x depends on the thermal history
and the preparation conditions. A systematic study of the

S.R. Naik et al. / Journal of Alloys and Compounds 566 (2013) 5461

signicance of the method used in the preparation of CoFe2O4


nanocrystals would provide insights towards better understanding
of the correlations between the particle dimension, spin orbit (LS)
coupling and the magnetic properties. Consequently, the desirable
magnetic properties for technological/biomedical applications can
be optimized through chemical changes which are sensitive to the
preparation issues such as method, reaction conditions, particle
size distribution and the occupancy of the two lattices by the metal
ions.
Solgel autocombustion method is among the best method and
used due to its effectiveness in obtaining the desired monophasic
products with greater homogeneity and ner particle size at very
low sintering temperatures [1618]. However to achieve the compositional homogeneity of the nal oxide powder, the preparation
of a homogenous gel with respect to the distribution of cations is
crucial. Therefore, it is essential to prepare a suitable precursor
solution which can be converted to a gel without any cation segregation. The usage of malic acid and ethylene glycol as complexing
and gelling agents at a neutral pH for the solgel assisted combustion synthesis of CoFe2O4 is reported herein. The effect of pH on
the chelating action of the organic acid is demonstrated, which is
vital in preventing the precipitation and aggregation of the ions
and maintaining the homogeneity of the solution [19]. A detailed
investigation about the inuence of distribution of the metal ions
(Co2+ and Fe3+) in the crystal lattice and the dimensions of the
ferrite oxide on the resultant magnetic properties is reported in this
article. The contribution of LS coupling, from Co2+ ions, towards
higher magnetic anisotropy and hence the magnetic properties is
investigated. The results provide an insight on the inter-relationship
of the particle dimension, the LS coupling and the resulting
superparamagnetic property. The conclusions on structural and
magnetic properties have been aptly supported by the magnetic
and spectral data obtained from various characterization techniques and the results prove CoFe2O4 nanoparticles to be a suitable
candidate in the technological and biomedical applications.
2. Experimental
2.1. Preparation
For the synthesis of CoFe2O4, stoichiometric amounts of analytical grade
Co(NO3)26H2O (SigmaAldrich) and Fe(NO3)39H2O (SigmaAldrich) were utilized
and were brought into solution by dissolving in double distilled water. The solution
was maintained around 100 C with continuous stirring. Calculated amount of
malic acid was than added. pH of the solution was then increased (around neutral)
with the slow addition of 30% ammonia solution. A considerable change in color
from light pink to dark wine red was observed which conrmed the chelating action
of malic acid. The dependence of chelating action on pH, seems to be very high as
there was no chelation and hence gel formation observed at lower pH values. It
can be said that the higher pH (around neutral), forces the organic acid to release
the acidic proton and in turn bind with the available metal ions. After conrming
the pH, ethylene glycol in the ratio 1:4 (with respect to the malic acid) was added.
The solution was then allowed to concentrate with continuous stirring and the gel
so obtained was heated in an oven at a temperature of about 200 C for 3 h. Formation of a voluminous foamy precursor was observed which was then crushed into
ne powder with the help of an agate mortar and pestle. The precursor was than
calcined at 400 C for 4 h. It was again ground with acetone and after drying was
sintered at 600 C for 6 h. The as-obtained powder was than subjected to various
characterization techniques.
2.2. Characterization
The crystallinity, crystal structure and the phase purity of the powders were
investigated by X-ray diffraction (XRD) technique using Cu Ka radiations of wavelength 1.5418 (ltered through Ni), in steps of 0.02 on a RIGAKU ULTIMA IV X-ray
diffractometer. SHIMADZU FTIR PRESTIGE-21 spectrophotometer was put into use
to record the FTIR spectrum of the oxide calcined at 600 C. A comparative study
of the Raman and IR actives modes was done. Raman spectrum was recorded in
the backscattering geometry in the range 1001000 cm1 using a HORIBA JOBIN
YVON HR-800 Raman spectrometer with an Olympus microscope (objective 50x)
attachment and equipped with a CCD detector. A 488 nm Ar+ ion laser with
10 mW powers was used as the excitation source for a spot of about 1 mm in diam-

55

eter. Morphology analysis of the compound was carried out on JEOL JSM-6360 LV
Scanning Electron Microscope (SEM). Transmission Electron Microscopy (TEM)
images were recorded on PHILIPS CM 200 Transmission Electron Microscope operating at an accelerating voltage of 200 kV and providing a resolution of 2.4 . Particle size distribution study of the as prepared nanoparticles was carried out by
employing a DELSA NANO S, BECKMAN COULTER, USA. The valence state and binding energies of various chemical species were determined by the X-ray photoelectron spectroscopy (XPS) employing VSW SCIENTIFIC INSTRUMENT with Al Ka as the
incident source having energy of 1486.6 eV with a resolution of 0.9 eV. The vaccum
maintained in the sample analyzer chamber was 1.4  108 Torr. 57Fe Mssbauer
measurements were performed in transmission mode with a 57Co radioactive
source in constant acceleration (triangular wave) mode using a standard PC based
Mssbauer spectrometer (Nucleonix Systems Pvt. Ltd., Hyderabad, India). Velocity
calibration of the spectrometer was done with an enriched a-57Fe metal foil absorber at room temperature and the isomer shift values were given relative to Fe foil.
The outer line width of calibration spectrum was 0.29 mm s1. The recorded
Mssbauer spectrum was tted by using the WinNormos t programme. A QUANTUM DESIGN PPMS-VSM magnetometer was used for magnetic characterization of
the pelletized compounds. The variation of the DC-susceptibility of each sample
with temperature was measured from 5 K to 300 K in the ZFC (zero eld cooling)
and FC (eld cooling) modes using a magnetic eld of 10 KOe. The magnetization
with varying magnetic eld of up to 50 KOe was also measured at 5 K and 300 K.

3. Results and discussion


3.1. Structural analysis
Fig. 1a shows the X-ray diffractogram of CoFe2O4 sintered at
600 C for 6 h. All the diffraction peaks observed for the oxide corresponds to the cubic spinel ferrite structure. The phase analysis
was carried out by matching the obtained diffractogram with the
standard ICDD card number 22-1086. The XRD pattern reveals
the monophasic formation of the polycrystalline compound and
also signies its greater crystallinity. A higher peak width which
is obvious from the XRD peaks signies that the compound comprises of ner particles. The broadening of the peaks also indicates
the decrease in the density and an increase in the surface to volume ratio of the CoFe2O4 nano oxide. The unit cell parameter (a)
is calculated for the CoFe2O4 system as a Voigt function for the
(3 1 1) peak. Crystallite size and the lattice strain present in the system are calculated from the X-ray diffractograms. WilliamsonHall
extrapolations as a Lorentzian function are very well utilized in calculating these parameters. WilliamsonHall plot for CoFe2O4 is
shown in Fig. 1b. A microdistortion of 2.7  102 is present in
the CoFe2O4 structure. This may be accompanied with the migration of Co2+ ions from the tetrahedral to the octahedral lattice. This
process of migration decreases the induced lattice strain, thereby
stabilizing the crystal structure [16]. The average crystallite size
and the lattice constant (a) obtained for the CoFe2O4 system from
the X-ray diffraction is 8 nm and 8.390 . The value for the lattice
constant is in agreement with the reported value [8].
3.2. Morphology and chemical analysis
The SEM micrograph of CoFe2O4 sintered at 600 C is presented
in Fig. 2. The image displays a aky and porous morphology for the
oxide ferrite. The micrograph conrms the formation of secondary
particles or lumps of aggregates which may be due to the agglomeration of the primary particles. Consistency in the ferrite phase
formation is seen for the sample which is equally supported by
the results obtained from XRD.
3.3. Particle size analysis
The TEM images of CoFe2O4 along with the electron diffraction
pattern are presented in Fig. 3ac. The images reveal the partially
fused nature of the hexagonal particles with an average particle
size of 40 nm. The variation in particle size is prominently seen
for the oxide. The well crystallized nature of CoFe2O4 along with

56

S.R. Naik et al. / Journal of Alloys and Compounds 566 (2013) 5461

Fig. 1. (a) X-ray diffraction pattern of cobalt ferrite nanoparticles sintered at 600 C for 6 h, and (b) WilliamsonHall plot of cobalt ferrite.

3.4. Raman spectroscopy

Fig. 2. SEM micrograph of cobalt ferrite heated at 600 C.

particle broadening is obvious from the electron diffraction pattern


from which the position of the individual planes for the spinel
CoFe2O4 can be clearly seen. This proves the calcination conditions
to be appropriate in yielding nanosized polycrystalline CoFe2O4
particles.

Raman spectroscopy is a highly sensitive tool for many lattice


effects, such as structure transition, lattice distortion, chargelattice and spin-lattice couplings, local cation distribution, and
magnetic ordering. CoFe2O4 has a cubic inverse/mixed ferrite
structure with O7h (Fd3m space group) which gives rise to 39 normal vibrational modes, out of which ve are Raman active.
Room-temperature Raman spectrum of CoFe2O4 nanoparticles in
the region 1001000 cm1 is presented in Fig. 4a. The spectrum reveals the single phase formation of the polycrystalline compound.
All the 5 (1 A1g + 1 Eg + 3 T2g) Raman active modes which are expected for the cubic spinel ferrite system are visible in the spectrum. The deconvoluted spectra (using suitable peak tting
software) are shown in Fig. 4bd. The spectra reveal the presence
of seven peaks in the Raman spectrum at 308, 357, 469, 577,
617, 658 and 693 cm1, most of which are the typical modes of
the cubic inverse-spinel ferrite structure. The T2g mode expected
at 190 cm1 for the compound is not clear from the spectrum,
but we do observe a peak at 308 cm1 which can be attributed to
the Eg mode [20]. The literature survey gives a precise idea about
the articles that reports a similar Raman spectrum for CoFe2O4
nanoparticles conrming our results [21]. Raman peak at
469 cm1 is related to the octahedral site mode that reects the
local lattice effect in the octahedral sublattice of CoFe2O4 [20].
The other intense peak at 693 cm1 (A1g mode) and a small peak

Fig. 3. Images depicting TEM (a) and (b), and electron diffraction pattern (c) of cobalt ferrite.

S.R. Naik et al. / Journal of Alloys and Compounds 566 (2013) 5461

57

Fig. 4. Graphs representing (a) comparison of the Raman and IR spectra for cobalt ferrite along with the deconvoluted patterns (b), (c) and (d).

at 617 cm1 are the symmetry vibrations of the metal in the tetrahedral site. A peak is observed in the region of 658 cm1 which can
be attributed to cobalt ions residing at defect sites that are probably located near the surface of the nanoparticles [20]. This peak is
pronounced due to the higher surface-to-volume ratio of the nanoparticles in this sample. Fe2O3 is the most common impurity phase
formed during the synthesis of spinel ferrites. It gives two sharp
signals at 240 cm1 and 300 cm1 with high intensity. The
240 cm1 peak is absent in the spectrum and a protrusion is observed at 308 cm1. This is comparatively weaker in intensity
and can therefore be said as the property of CoFe2O4 and not
Fe2O3 [22]. All together there are two different lattices in the spinel
ferrite structure, and an attempt is made over here to assign the
modes for the various signals observed in the spectrum based on
the idea obtained from the literature.
The difference between Raman and IR vibrational modes for
CoFe2O4 in the range 2751000 cm1 is displayed in Fig. 4a. There
are four different modes (m1, m2, m3 and m4) assigned for the tetrahedral lattice. All the four modes are Raman active while only 2 (m3
and m4) are IR active. From the gure, IR stretching modes are evident for the compound at 400 cm1 and 574 cm1, which are the
FeO symmetric and asymmetric stretching vibrations, along with
2 other signals at 279 cm1 and 302 cm1. On the contrary the signal evident at 393 cm1 which is highly intense in the IR region is
absent in the Raman spectrum proving it to be either a m3 or a m4
mode which are strong IR active modes and are weakly observed
in the Raman spectrum. A weakly intense signal in the Raman
spectrum is observed at 574 cm1 (asymmetric FeO stretching)
indicating that m3 and m4 must be strong IR active modes only.

3.5. XPS analysis


XPS is a powerful tool in determining the oxidation state and
binding energy of the various chemical species present in a compound. A change in the oxidation state brings along a change in
the number of unpaired electrons thereby changing the properties
of a system. The spectra of all the elements are deconvulated using
peak tting software. The full scan (FS) spectrum is presented in
Fig. 5a. From the gure, we can clearly observe the respective
peaks for O 1s, Fe 2p and Co 2p which conrms the elements to
be present in the system. We could verify the absence of carbon
in the material based on the absence of the carbon peak which normally originates at 285 eV. The valence state of iron and cobalt decides the overall magnetic moment of the system. The compound
under study is prepared in an oxygen rich atmosphere and therefore there is possibility of redox reaction to occur in which electron
exchange can take place between Co2+ and Fe3+ to form Fe2+ and
Co3+ ions. This might bring along a change in the overall magnetic
moment. XPS spectrum of Fe 2p is presented in Fig. 5b. The two
signals at 725.46 eV and 711.94 eV represent the Fe 2p1/2 and Fe
2p3/2 for the Fe3+ state, conrming the valence state of iron to be
Fe3+and not Fe2+. This also rules out the possibility of metal iron
to be present. The corresponding spectrum for Co 2p is presented
in Fig. 5c. From the gure, signal for Co 2p1/2 and Co 2p3/2 are observed at 795 eV and 779 eV respectively which are characteristic
binding energies for the Co2+ ion. The binding energy of the oxygen
having value of 529.4 eV corresponding to the O 1s is presented in
Fig. 5e. Another peak at 531.6 eV is also observed which might be
assigned to the defect oxide or the surface oxygen ions with low

58

S.R. Naik et al. / Journal of Alloys and Compounds 566 (2013) 5461

Fig. 5. Deconvulated XPS spectra showing (a) full scan (b) Fe 2p (c) Co 2p and (d) O 1s of cobalt ferrite.

coordination situation [23,24].The values obtained for the binding


energies of all the elements are in agreement with the reported
ones [25]. Studies conrm the existence of the metal ions in the required valency thereby enhancing the overall magnetic moment.

3.6. Mssbauer spectroscopy


Apart from particle dimensions of the compound and valence
state of the metal ions, magnetic properties are also governed by
the distribution of the metal ions in both the sublattices (tetrahedral and octahedral) of spinel oxides. The degree of interaction between metal ions is fully dependent on the type of metal ions that
occupy these two sublattices. As discussed earlier, an increase or
decrease in the magnetic property is a result of such interaction
and therefore it is necessary to probe this distribution of metal
ions. Co2+ ion has the electronic conguration of 3d7. In CoFe2O4,
Co2+ preferentially enters the octahedral site. Since the oxygen
atoms surrounding Co2+ ion exert a weak eld environment, the
degenerate d-orbitals of Co2+ splits to give the resultant high spin
electronic conguration of t 52g e2g . With this electronic conguration, Co2+ ion possesses the crystal eld stabilization energy of
0.8 Do and preferentially occupies the octahedral site [26,27]. Reports in the literature reect on the random nature of CoFe2O4 and
therefore application of Mssbauer spectroscopy becomes necessary to investigate the degree of inversion [8,11,2830]. Mssbauer
spectrum of CoFe2O4 sample, recorded at room temperature is presented in Fig. 6. The presence of Mssbauer nuclei in ve different
chemical environments is conrmed from the ve six lines hyperne patterns (Zeeman splitting or sextet) shown in the gure. This

also conrms the ferrimagnetic nature of the compounds. The


Mssbauer spectrum is not showing any superparamagnetic
behavior at room temperature.
Mssbauer parameters obtained after tting the spectrum are
presented in Table 1. The values obtained for the isomer shifts
(d) of all the ve sextets are 0.206 mm s1 for tetrahedral and
0.3310.377 mm s1 for octahedral site respectively, which conrm the presence of iron in the Fe3+ high spin state [31]. These results are in compliance with the ones obtained from the XPS study.

Fig. 6. Room temperature Mssbauer spectrum of cobalt ferrite.

S.R. Naik et al. / Journal of Alloys and Compounds 566 (2013) 5461

59

Table 1
Mssbauer tting parameters of CoFe2O4 obtained using Win NORMOS tting
programme. Isomer shift (d), quadrupole splitting (D), line width (C), hyperne
magnetic eld (Hhf) and relative areas (RA) are shown in the table. All the results are
relative to Fe metal foil.
Fe sites
Sextet
Sextet
Sextet
Sextet
Sextet

1
2
3
4
5

(tet)
(oct)
(oct)
(oct)
(oct)

d (mm s1)

D (mm s1)

C (mm s1)

Hhf (T)

RA (%)

0.206
0.377
0.331
0.349
0.346

0.019
0.010
0.071
0.061
0.006

0.232
0.391
0.148
0.394
0.192

48.59
51.29
46.78
43.58
48.74

22.63
23.17
12.53
13.54
28.13

Isomer shift values are in the range of 0.900.95 mm s1 for tetrahedral and 1.051.10 mm s1 for octahedral site, respectively for
Fe2+ ions [3234]. So the hyperne structure does not evidence
of the presence of Fe2+ ions. The quadrupole splitting (D) values
are found to be low indicating the overall presence of cubic symmetry at both the sites. The magnetic hyperne eld results conrm the distribution of the Mssbauer nuclei in all the ve
sublattices (four octahedral and one tetrahedral) of the compound.
The magnetic hyperne elds for a tetrahedral and four octahedral
sublattices are 48.59, 51.29, 48.74, 46.78, and 43.58 Tesla respectively. The A and B sites were assigned to the sextets based on isomer shift and the magnetic hyperne eld values as reported
[31,35,36]. Since the bond separation between Fe3+O2 is larger
for octahedral ions as compared to that of the tetrahedral ions,
smaller overlapping of the orbitals of Fe3+ ions and oxygen anions
(low s electron density compared to tetrahedral site) and the smaller covalency lead to larger isomer shift at the octahedral site. So
the sextet with a lower isomer shift corresponds to the tetrahedral
site, and the sextet having a higher isomer shift represents the
octahedral sites [35]. The percent distribution of the Fe3+ in the tetrahedral and the octahedral sublattices is calculated from the relative areas of the hyperne sextets. The results display the
distribution to be 22.6% (tetrahedral site) and 77.4% (octahedral
site), respectively showing the degree of inversion to be 0.45 with
the formula (Fe0.45Co0.55)[Co0.45Fe1.55]O4. This proves the possibility of the cation exchange taking place in the nanocrystalline
CoFe2O4 reported herein (CoFe2O4 is an inverse spinel in bulk
form).

3.7. Magnetic measurements


The dependency of the magnetic property on temperature is reported herein. The ZFC and FC susceptibility curves of CoFe2O4
nanoparticles are given in Fig. 7. From the gure, a net irreversibility is observed between the ZFC and FC curves i.e. the ZFC curve
shows a clear maximum at a critical temperature and decreases
rapidly to zero with decreasing temperature, while FC remains almost constant below this critical temperature, which is typical of a
superparamagnetic behavior. The as-synthesized nanoparticles can
therefore be considered as magnetic single domains with a blocking temperature (TB) of 235 K, corresponding to the maximum value of the ZFC curve.
It is well known that CoFe2O4 can deviate from the thermodynamically stable structure and can be described by the appropriate
2
3
3
formula as (Co2
k Fe1k )Co1k Fe1k O4 , where parentheses and
square brackets correspond to the tetrahedral and octahedral lattices of the spinel structure, respectively. The degree of structure
inversion is measured by k < 1. For instance, k is found to be about
0.4 [8] and 0.84 [11] in 8 nm and 5.5 nm sized CF particles respectively, prepared by the coprecipitation method and polyol method.
It is estimated to be about 0.69 [28] and 0.8 [29] in 20 nm and 5 nm
sized particles respectively prepared by the micellar technique. In
all cases, the magnetic properties appear to be sensitive to the cat-

Fig. 7. Variation of magnetization with eld cooling (FC) and zero eld cooling
(ZFC) from 5 to 300 K. A magnetic eld of 10 kOe is maintained for the FC
measurement.

ion distribution between the octahedral and tetrahedral sites. A


certain number of Co2+ cations migrate from the octahedral sites
to the tetrahedral ones, accompanied by a reverse transfer of Fe3+
ions from the tetrahedral to the octahedral ones in order to relieve
the strain. The population of Co2+ cations in the octahedral spinel
sites decreases, which affects the magnetic properties of CoFe2O4
by lowering the single ion anisotropy of Co2+ ions present in the
crystal lattice and thereby affecting the TB value. Mssbauer spectroscopy is employed to conrm the structural deviation from an
inverse-spinel-like lattice of the particles reported herein. The degree of inversion is found to be 0.45.
The major factor for determining the superparamagnetic properties of nanocrystals is magnetocrystalline anisotropy. It originates from the LS coupling at the crystal lattices. According to
StonerWohlfarth single domain theory [30], the magnetocrystalline anisotropy energy, EA, of a single domain nanocrystal is
approximated by EA = KV sin2h where K is the magnetocrystalline
anisotropy constant, V is the volume of nanocrystal, and h is the angle between the easy axis of nanocrystal and the direction of eldinduced magnetization. The magnetocrystalline anisotropy serves
as an energy barrier to block the spin relaxation, which changes
the magnetic state from ferrimagnetic to superparamagnetic. The
height of EA determines the blocking temperature at which the
thermal activation can overcome EA and the nanocrystals transform into the superparamagnetic state. Both the magnetocrystalline anisotropy constant K and the volume of nanocrystals
control the magnetic anisotropy. The magnitude of K is closely related to the strength of LS coupling. As a result, the superparamagnetic properties of nanocrystals can be directly correlated
with the variation of LS coupling strength.
The magnetic anisotropy is the result of the accumulative contribution of individual magnetic cations. Therefore, the magnetic
coupling at the individual lattice site can be straightforwardly correlated to magnetic properties such as superparamagnetism. For a
crystal with a cubic spinel structure, metal cations occupy the tetrahedral (A site) and octahedral (B site) lattice sites. Since the ligand eld is weak in spinel ferrites, all cations assume high spin
states. A Fe3+ cation with 3d5 electronic conguration usually has
its orbital angular momentum quenched in a weak ligand eld.
Therefore, the contribution to the magnetic anisotropy should only
come from Co2+ cations in CoFe2O4. The LS coupling strength of
the Co2+ determines the relative magnitude of magnetic anisotropy. The cubic symmetry is often lowered to a trigonal eld in spinels due to the structural distortion from the JahnTeller effect
and/or the nonstructural distortion from the interactions between

60

S.R. Naik et al. / Journal of Alloys and Compounds 566 (2013) 5461

4. Conclusion

Fig. 8. Graphs depicting the variation in the magnetic properties of the cobalt
ferrite nanoparticles at 5 K and 300 K in the magnetic eld of 50 kOe.

the central cation and the cations outside the nearest neighborhood of the central ion. A Co2+ cation with 3d7 electronic conguration at B site in CoFe2O4 has a triplet 4T1g ground state. Even
though the trigonal eld is introduced with the T1g ground state
further splitting into A2 and E states, the Co2+ cation with a degenerated ground state of E is still considered to have a strong LS coupling, and therefore contributes greatly to the magnetic anisotropy
of CoFe2O4 [13]. The strong LS couplings at Co2+ lattice sites generate a large anisotropy constant K and results in much higher
anisotropy energy barrier in CoFe2O4 nanocrystals. Hence, we observe higher blocking temperature for CoFe2O4 nanocrystals.
The coercivity of magnetic nanocrystals is surely related to the
magnetic anisotropy. At a given temperature, the required magnetic eld strength for overcoming anisotropy to ip the magnetic
spin increases with increasing anisotropy energy barrier. Strong L
S couplings in Co2+ cations result in much higher coercivity in
CoFe2O4. The coercivity of nanocrystals also has a contribution
from the surface anisotropy of nanocrystals. Since the compound
under study consist of the particles of varying sizes, it is very difcult to gauge the inuence of the particle size on the resultant
coercivity.
The display of superparamagnetic properties by CoFe2O4 nanocrystals facilitates the fundamental understanding of magnetism at
the nanometer scale. The studies conrm that CoFe2O4 is a better
magnetic material than magnetite. The presence of Co2+ ions in
the octahedral lattice and the contribution of LS coupling towards
the resultant magnetic properties are signicantly higher. The LS
contribution of Fe2+ ions in magnetite is lower than Co2+ in CoFe2O4
and therefore CoFe2O4 nanocrystals clearly display better magnetic
characteristics than Fe3O4 nanocrystals. Clearly, the CoFe2O4 nanoparticles are good candidates for technological applications in
terms of magnetic characteristics.
Fig. 8 represents the dependence of magnetization of CoFe2O4
nanoparticles on magnetic eld at 300 K and 5 K. At 300 K these
nanoparticles exhibit hysteresis loops with coercivity, remanance
and a saturation magnetization of 1326 Oe, 33 emu g1 and
74 emu g1 respectively. At 5 K the saturation magnetization value
increases to 80 emu g1. This increase in the values may be because of the proper ordering of the magnetic moments along the
direction of the eld, which in turn are randomized at elevated
temperatures (at 300 K) due to the thermal vibrations. The coercivity and remanance are found to be 13939 Oe and 68 emu g1
respectively. The squareness ratio (Mr/Ms) is found to increase from
0.44 (300 K) to 0.86 (5 K).

Preparation of the CoFe2O4 nanocrystalline compound was carried out successfully by the solgel autocombustion method. The
structural characterization like XRD proved the monophasic formation and conrmed the nanocrystalline nature of the particles.
Partial migration of the cobalt ions was suspected from the microdistortion calculated using the WilliamsonHall plots. This was
conrmed from the Mssbauer spectrum which showed the degree
of inversion to be 0.45 and the values for isomer shift veried the
existence of iron in the Fe3+ high spin state. XPS conrmed the
presence of the metal ions in the required valence state, which enhanced the overall magnetic moment. Raman spectroscopy effectively demonstrated the formation of the spinel phase only
thereby eliminating the probability of the formation of non-magnetic a-Fe2O3 phase. The LS contribution from Co2+ ions in
enhancing the magnetic properties was effectively demonstrated.
A large variation in the magnetic properties was observed with
varying eld and temperature. An increase in the value of saturation magnetization was observed at 5 K as compared to 300 K, signifying a decrease in the randomization of the magnetic moments
by the thermal vibrations. The acquisition of lattices by the metal
ions due to a probable migration was effectively proved by the
Mssbauer spectroscopy which conrmed the formation of the
random cobalt ferrite, thereby supporting the explanation provided for the magnetic properties exhibited by the material under
study.
Acknowledgement
S.R. Naik and A.V. Salker would like to acknowledge Dr. A.
Banerjee, Dr. V. Sathe, Dr. T. Shripathi and Dr. S. Tripathi (UGCDAE Consortium for Scientic Research, Indore, India) for providing
VSM, Raman and XPS facility. Dr. Rahul Mohan and Ms. Sahina Gazi
(NCAOR, Goa, India) are acknowledged for providing SEM facility.
References
[1] Q. Song, Z.J. Zhang, J. Am. Chem. Soc. 126 (2004) 6164.
[2] L.D. Tung, V. Kolesnichenko, D. Caruntu, N.H. Chou, C.J. OConnor, L. Spinu, J.
Appl. Phys. 93 (2003) 7486.
[3] T. Hyeon, Y. Chung, J. Park, S.S. Lee, Y.-W. Kim, B.H. Park, J. Phys. Chem. B. 106
(2002) 6831.
[4] R.S. Devan, Y.D. Kolekar, B.K. Chougule, J. Phys. Condens. Matter 18 (2006)
9809.
[5] N. Sivakumar, A. Narayanasamy, K. Shinoda, C.N. Chinnasamy, B. Jeyadevan, J.M. Greneche, J. Appl. Phys. 102 (2007) 0139161.
[6] B.X. Gu, Appl. Phys. Lett. 82 (2003) 3708.
[7] F. Cheng, C. Liao, J. Kuang, Z. Xu, C. Yan, L. Chen, H. Zhao, Z. Liu, J. Appl. Phys. 85
(1999) 2782.
[8] M.A.G. Soler, E.C.D. Lima, S.W. da Silva, T.F.O. Melo, A.C.M. Pimenta, J.P.
Sinnecker, R.B. Azevedo, V.K. Garg, A.C. Oliveira, M.A. Novak, P.C. Morais,
Langmuir 23 (2007) 9611.
[9] N. Moumen, M.P. Pileni, Chem. Mater. 8 (1996) 1128.
[10] E. Manova, B. Kunev, D. Paneva, I. Mitov, L. Petrov, C. Estourns, C. DOrlans, J.L. Rehspringer, M. Kurmoo, Chem. Mater. 16 (2004) 5689.
[11] S. Ammar, A. Helfen, N. Jouini, F. Fivet, I. Rosenman, F. Villain, P. Molini, M.
Danot, J. Mater. Chem. 11 (2001) 186.
[12] E.V. Groman, J.C. Bouchard, C.P. Reinhardt, D.E. Vaccaro, Bioconjugate Chem.
18 (2007) 1763.
[13] S. Rana, J. Rawat, M. Sorensson, R.D.K. Misra, Acta Biomater. 2 (2006) 421.
[14] S. Rana, A. Gallo, R.S. Shrivastava, R.D.K. Misra, Acta Biomater. 3 (2007) 233.
[15] J. Rawat, S. Rana, M.M. Sorensson, R.D.K. Misra, Mater. Sci. Technol. 23 (2007)
97.
[16] S.R. Naik, A.V. Salker, J. Mater. Chem. 22 (2012) 2740.
[17] S.R. Naik, A.V. Salker, Phys. Chem. Chem. Phys. 14 (2012) 10032.
[18] G. Ji, S. Tang, B. Xu, B. Gu, Y. Du, Chem. Phys. Lett. 379 (2003) 484.
[19] P. Vaqueiro, M.A. Lpez-Quintela, Chem. Mater. 9 (1997) 2836.
[20] G. Shemer, E. Tirosh, T. Livneh, G. Markovich, J. Phys. Chem. C 111 (2007)
14334.
[21] M.A.G. Solera, T.F.O. Melo, S.W. da Silva, E.C.D. Lima, A.C.M. Pimenta, V.K. Garg,
A.C. Oliveira, P.C. Morais, J. Magn. Magn. Mater. 272276 (2006) 2357.
[22] L.B. Tahar, L.S. Smiri, M. Artus, A.-L. Joudrier, F. Herbst, M.J. Vaulay, S. Ammar,
F. Fivet, Mater. Res. Bull. 42 (2007) 1888.

S.R. Naik et al. / Journal of Alloys and Compounds 566 (2013) 5461
[23] S. Liang, F. Teng, G. Bulgan, R. Zong, Y. Zhu, J. Phys. Chem. C 112 (2008)
5307.
[24] M. Liu, G.J. Zhang, Z.R. Shen, P.C. Sun, D.T. Ding, T.H. Chen, Solid State Sci. 11
(2009) 118.
[25] G.B. Ji, S.L. Tang, S.K. Ren, F.M. Zhang, B.X. Gu, Y.W. Du, J. Cryst. Growth 270
(2004) 156.
[26] H.S.T.C. ONeil, A. Navrotsky, Am. Mineral. 68 (1983) 181.
[27] J.K. Burdett, G.D. Price, S.L. Price, J. Am. Chem. Soc. 104 (1982) 92.
[28] C. Liu, B. Zou, A.J. Rondinone, Z.J. Zhang, J. Am. Chem. Soc. 122 (2000) 6263.
[29] N. Moumen, P. Bonville, M.P. Pileni, J. Phys. Chem. 100 (1996) 14410.
[30] C.R. Vestal, Z.J. Zhang, Nano Lett. 3 (2003) 1739.

61

[31] A.M. Banerjee, M.R. Pai, S.S. Meena, A.K. Tripathi, S.R. Bharadwaj, Int. J.
Hydrogen Energy 36 (2011) 4768.
[32] K. Sharma, S.S. Meena, S. Saxena, S.M. Yusuf, A. Srinivasan, G.P. Kothiyal, Mater.
Chem. Phys. 133 (2012) 144.
[33] K. Sharma, S.S. Meena, C.L. Prajapat, S. Bhattacharya, Jagannath, M.R. Singh,
S.M. Yusuf, G.P. Kothiyal, J. Magn. Magn. Mater. 321 (2009) 3821.
[34] K. Sharma, A. Dixit, S.S. Meena, Jagannath, S. Bhattacharya, C.L. Prajapat, P.K.
Sharma, S.M. Yusuf, A.K. Tyagi, G.P. Kothiyal, Mater. Sci. Eng. C 29 (2009) 2226.
[35] S.S. Shinde, S.S. Meena, S.M. Yusuf, K.Y. Rajpure, J. Phys. Chem. C 115 (2011)
3731.
[36] D.H. Lee, H.S. Kim, C.H. Yo, K. Ahn, K.H. Kim, Mater. Chem. Phys. 57 (1998) 169.

You might also like