You are on page 1of 10

Journal of Alloys and Compounds 653 (2015) 513e522

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Surface and interparticle interactions effects on nano-cobalt ferrites


M. Saidani a, *, W. Belkacem a, A. Bezergheanu b, C.B. Cizmas b, N. Mliki a
a
b

LMOP: LR99ES17, Facult


e des Sciences de Tunis, Universit
e de Tunis El Manar, 2092, Tunisia
D
epartement d'Ing
enierie Electrique et Physique Appliqu
e, Universit
e Transilvania de Brasov, Romania

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 1 August 2015
Accepted 1 September 2015
Available online 7 September 2015

Cobalt ferrites nanoparticles CoxFe3-xO4 (1 x < 1.8) were synthesized by a solvothermal route. X-ray
diffraction, Transmission Electron Microscopy and magnetic measurements show a strong correlation
between the structural and the magnetic properties. The highest cobalt amount substituted samples
(x 1.6 and 1.8) show a jump at low temperature in the hysteresis loops in the range of low applied elds
(<10 kOe). This behavior has been explained by surface effects and interparticle interactions. A difference
has been found using two different estimation of the magnetic anisotropy: from the law of approach to
saturation and using the Neel-Brown approach. The last approach leads to an increase of the anisotropy
for x 1.6 and 1.8 coming from nite size effects. Surface effects have been pointed out using the Bloch
law tting. Zero Filed Cooling (ZFC) measurements exhibit a blocking temperature (TB) range from 294 to
240 K. Kneller's law tting of the coercive eld dependence on the temperature exhibits a non-neglected
interparticle interaction and a wide size distribution according to the difference found between TB and
the Kneller transition temperature TBK.
2015 Published by Elsevier B.V.

Keywords:
Nano-cobalt ferrites
Surface affects
Interparticle interaction

1. Introduction
Cobalt ferrite nanoscale materials have attracted a great
attention due to the continuous discover of new promising properties which differ from the bulk ones. Reduced sizes may present
interparticle interactions [1] and surface effects [2]. Cobalt Ferrites
nanoparticles are used in a wide variety of applications such as
recording media, microwave, biomedicine, photo-magnetism [36]. This can be related to their high cubic magnetic anisotropy,
their chemical stability and their high coercive eld [7]. These
samples have been synthesized by a several chemical and physical
routes in the purposes to nd the suitable one for the required
properties.
Cobalt ferrites, of the Spinel structure with AB2 O4 formula, can
3 Td 2 3 Oh
be described as A2
1t Bt At B2t O4 where Td , Oh refer to the
tetrahedral and octahedral sites respectively and t is the inversion
degree. It is called a normal (inverse) spinel if t is equal to 1 (0).
Accordingly, the magnetite Fe3O4 crystallizes in a completely inverse spinel structure. Fe3 , occupying the octahedral and tetrahedral sites by half, are antiferromagnetically coupled through the
magnetic superexchange interaction. Fe3 and Fe2 cations, located

* Corresponding author.
E-mail address: mohamed1saidani@gmail.com (M. Saidani).
http://dx.doi.org/10.1016/j.jallcom.2015.09.020
0925-8388/ 2015 Published by Elsevier B.V.

in the octahedral sites, are ferromagnetically coupled by means of


the magnetic double exchange interaction [8]. Thus, the magnitude
of those interactions depends on the cationic distribution and the
type of the tetrahedral and octahedral cations. The substitution of
Fe2 by Co2 affects the physical properties regarding to the
dimension and the difference between their magnetic natures.
Swatzky et al found that Fe-OCo superexchange interaction is
weaker than FeeCoFe [9].
In this work, we are interested in the effect induced by the
substitution the iron by cobalt. It is helpful to have knowledge on its
real structure and on what it depends. CoFe2O4 has an inverse
spinel structure. Although, it was found that a ratio of Co2 (2e24%)
can occupy tetrahedral sites [10]. This is strongly dependent on the
synthesis method and the heat treatment [11]. Thus, results are
governed by many parameters as mentioned above and, from a
synthesis technique to another, strange and interesting effects
could be observed. The grain size, also, is one of the principal parameters that affects the magnetic properties and was found by
Ref. [2] to be in relationship with the temperature. In addition, the
substituted amount has a non-neglected effect on the grain size
giving rise to different magnetic features [12].
Here, an interesting effect on the hysteresis loops is discussed in
terms of surface and interparticle interactions resulting from an
increase of the cobalt amount. We investigate, to our knowledge,
for the rst time the effect of a high cobalt amount substituted iron

514

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

on cobalt ferrite using this synthetic technique on the magnetic


properties.
2. Experimental procedure
Samples of CoxFe3-xO4 (x 1 (S1), 1.2 (S2), 1.4 (S3), 1.6 (S4), 1.8
(S5)) were synthesized by a solvothermal chemical route [13]. All
samples were characterized by X-ray diffraction (XRD) using a
Siemens D5000 diffractometerworking with Cu Ka radiation in the
range of Bragg angles between 20 and 80 . The scanning rate is of
0.033 at room temperature. Morphology, grain size and chemical
compositions of the nanoparticles were determined by Transmission Electron Microscopy (TEM), coupled with EDS, using a
JEOL2010 FEG microscope operating at 200 kV. Magnetic measurements were carried out using a Vibrating Sample Magnetometer (VSM) at several temperatures ranging from 10 to 300 K. Zero
eld cooling (ZFC) curves were taken at an applied eld of 100 Oe
from 40 to 320 K.
3. Structural and magnetic results
3.1. Structural characterization
Fig. 1 shows X-ray diffraction patterns acquired for all samples
with different cobalt contents which could be indexed in a spinel

cubic structure. No addition peaks were detected proving a good


homogeneity of elaborated samples. The lattice parameter was
extracted from Rietveld renement method of the diagrams using
FullProf software. An example of the tted one is given in Fig. 1b for
x 1 c2 1:14. Fig. 2 displays the dependence of the lattice
parameter on the cobalt content. S1 (CoFe2O4) was found to have a
lattice parameter of about 8.4 which matches the value reported
in the literature [13]. As the cobalt amount increases, the lattice
parameter has been found to slightly decrease.
It is worthy to note that when a solid substitution takes place,
the lattice parameter follows the known Vegard's law [14,15]. This
is due to the difference between the atomic radius of cobalt and
iron. Nevertheless, a monotonic decrease of the lattice parameter is
observed in this work. This behavior hasto be taken into consideration. First of all, it is a deviation from the Vegard's law as it has
been observed in several works [16], which could be due to several
effects. For spinel ferrites, the cell parameter can be calculated
using the formula [17]:


p
8 
atheoretic p  rA R0 3rB R0
3 3

(1)

where rA , rB and R0 are the tetrahedral, the octahedral and the


oxygen radius respectively. As it can be seen from the equation
above, the lattice parameter is very sensitive to the cationic distribution. Furthermore, the oxidation state is also an important
parameter that cannot be neglected regarding the fact that
Fe3 Co3 has a less radius than Fe2 Co2 .
As it has been mentioned above, an amount of Co2 (till 24%) can
be located in the tetrahedral sites. When x exceeds 1, this ratio can
vary, an oxidation of Co2 to Co3 [18] can occur and/or a nonstoichiometry caused by the absence of some oxygen anions as in
MgxFe3-xO4 [19]. One concludes that so many important factors are
responsible for determining the structural details.
Table 1 lists the EDS analysis of the cobalt content. A deviation
from the nominal values for S4 and S5 has been found. This may
come from supersaturating or other parameters which still not well
understood. Fig. 3 illustrates TEM images corresponding to S1, S3,
S4 and S5 samples. From these images, one notes that the nanoparticles have almost spheroid shape. Fig. 4 displays the crystallite
and grain size, estimated using the Scherer's formula and TEM
distributions analysis (see Fig. 5) respectively. It can be seen that
the sizes are fairly dependent on the cobalt content. S4 and S5 have
been found to present the smallest sizes. Note that, the

Lattice constant ( )

8,42

8,40

8,38

1,0
Fig. 1. XRD patterns for CoxFe3-xO4 samples (a) and an example of tted diagram for
x 1 (b).

1,2

1,4

cobalt content x

1,6

Fig. 2. Cell parameter as a function of the cobalt content x.

1,8

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522


Table 1
Mean composition of the CoxFe3-xO4nanoparticles.
Labels

x nominal

x measured by EDS analysis

S1
S2
S3
S4
S5

1
1.20
1.40
1.60
1.80

1
1.19
1.32
1.34
1.50

nanoparticles for S4 (see Fig. 3) are rather agglomerated when


compared to the other samples.
The nanoparticles have a spheroid shape except for x 1.8
where particles have an irregular shape.
Regarding the low sizes of the particles, the spin reversal dynamics will be directly affected because both exchange and dipolar
interparticle interactions will be affected [20].
4. Magnetic measurements
The sizes of the nanoparticles are less than 10 nm (see Fig. 4).
These sizes may present the superparamagnetic behavior [2]. We
have performed Zero Field Cooling (ZFC) measurements in the
temperature range between 10 and 320 K under an applied eld of
100 Oe. Fig. 6 shows the measurements for all cobalt contents. Note
that as the cobalt content increases the maximum of the peaks

515

become more and more broad. This result provides a wide particle
size distribution as conrmed also by TEM analysis.
Fig. 7aee display hysteresis loops for the whole samples recorded at 10, 50, 150, 250 and 300 K. As the temperature and the cobalt
content increase, hysteresis shape and parameters (coercive eld:
Hc, saturation magnetization: Ms and remanence magnetization:
Mr) exhibit a noticeable variations. At 300 K (Fig. 7d), CoFe2 O4 has
the typically hysteresis loops behavior for soft ferrimagnetic materials with a coercive eld of 263 Oe comparable to the value
found by Zhang et al. [21] for comparable sizes and a saturation
magnetization Ms 79 emu/g, close to the bulk value (80 emu/g)
[17]. The other samples present the superparamagnetism behavior.
At 10 K (Fig. 7a), S1, S2 and S3 behave as a ferrimagnetic systems
with an increase of the coercive eld due to the decrease of thermal
agitation [21]. The samples (S4 and S5) with the highest cobalt
amount, exhibit a strange dependence of the magnetization on the
applied magnetic eld at low temperatures (Fig. 7b and d); a jump
has been occurred in the eld-range from 0 to 0.45 kOe.

5. Discussion
5.1. Hysteresis shape
The behavior (see Fig. 7) found on the hysteresis loops can be
related to different effects. Raghunathan et al [22] have found a

Fig. 3. TEM images for CoxFe3-xO4 (x 1: a, x 1.4: b, x 1.6: c and x 1.8: d)

516

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

10

x=1
x=1,2
x=1,4
x=1,6
x=1,8
0,05 CoxFe3-xO4

Moment (emu)

XRD, sherrrer formula


TEM

Size (nm)

6
0,00
40

4
1,0

1,2

1,4

cobalt content x

1,6

80

120

160

200

240

280

Temperature (K)

1,8

320

Fig. 6. ZFC measurements for samples CoxFe3-xO4.

Fig. 4. Crystallite and grain sizes as a function of the cobalt content (from TEM and
XRD measurements).

number of nanoparticles

similar behavior by applying a compressive stress on electrical


steels. Antunes et al [23] have found this behavior in Er(Co, Mn)O3
perovskite systems by a systematic substitution. Using the same
chemical synthesis route and changing the temperature, C. VazquezeVazquez et al [2] found this kind of jump. They associated it
to surface contribution regarding the sizes of their nanoparticles.Nevertheless, Zysler et al. [24] have studied the interparticle interactions effect in Fe0:2 Ni0:74 50 B50 and they showed that it

80

(a)

80

40

40

20

20

0
4

numbre of nanoparticules

(b)

60

60

80

strongly affects the shape of the hysteresis loop at low temperature.


Zhang et al. [21] observed at 10 K for CoFe2O4 a drop in the hysteresis loop nearby 0 applied elds and they related it to the existence of soft phases. However, their Kneller's law t showed the
existence of interparticle interaction which can contribute to the
drop as well as additional soft phases.
Note that the drop observed in the work of Zhang et al [21] has
been observed for S2 in our work but with a less magnitude. In
terms of surface anisotropy magnitude, Kachkachi et al [25] discussed how the strength of this jump would be.

size (nm) (x=1)

10

12

(c)

10

size (nm) (x=1,4)

11

12

(d)
40

60

40

20
20

0
4

size (nm) (x=1,6)

10

0
4

Fig. 5. Size distributions for different Co contents.

size (nm) (x = 1,8)

10

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

517

Fig. 7. Hysteresis loops: (a) for S1, S2 and S3 at 10 K (b) for S4 and S5 at 10 K. (c) for S5 at 50, 150 and 250 K. (d) for S1-5 at 300 K and (e) a zoom of S2 (x 1.2) at 10 K.

According to the results reported in the literature, our magnetic


results with the structural results as TEM images, the low sizes and
the HRTEM analysis for S5 both, interparticle interactions and
surface effects could be the cause of the behavior found for S4 and
S5. Furthermore, S4 and S5 present the lowest sizes: 4.8 and 6.2 nm
respectively. These sizes are small enough to exhibit an important
misalignment surface spin layer [26] which will have an important
role on the spins reversal dynamic. Comparing hysteresis loops of
S4 and S5, one can conclude that the effect strength is different. As
the temperature increases, these effects vanish at temperatures
higher than 100 K (see Fig. 7c).

6. Saturation magnetization and magnetic moment


In this part, the saturation magnetization MS, the magnetocristalline anisotropy deduced from MS, the effective anisotropy

deduced from ZFC measurements are discussed. A comparative


study will be done between the two kinds of anisotropy in order to
accentuate surface effects.
Even at high applied magnetic eld (60 kOe), hysteresis loops
don't exhibit any saturation. It is a direct consequence of a spin
misalignment at the surface and reects also that a fraction of the
nanoparticles still in a relaxation state even at low temperature
[27]. Nevertheless, the magnetization dependence on the magnetic
eld tends to saturate when the temperature increases (Fig. 6c). The
saturation magnetization Ms can be tted using the high eld part
of M (H) chosen between 0.99 Mmax and Mmax using the following
equation [28]:



b
MH Ms 1  2
H
H is the applied magnetic eld and b is a tting constant.

(2)

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

Fig. 8 displays the dependence of Mson the cobalt content x.


Fig. 9 (symbols) shows the dependence of Msvs the temperature.
Ms has been found to decrease with increasing the cobalt content x
as well as increasing the temperature for each sample. At 10 K, the
highest and lowest values of Ms were 102 and 49 emu/g for compositions x 1 and 1.8 respectively (Fig. 8). However, it is noticeable that as x increases the decrease of the Ms is getting to be faster.
This could be related to the decrease of the particle size [29], the
cationic distribution [30] and the oxidation state, Co3 being
diamagnetic [31]. Indeed, according to Neel's law, the magnetization of ferrites by formula unit can be calculated using the relation
given by Ref. [17]:

symbols : exp-data
lines : Bloch's law fit
90

MS (emu/g)

518

x=1
x = 1,2
x=1,4

60

x = 1,6
x = 1,8

m mB  mA

(3)

30
0

where mB and mA are the magnetization of the octahedral B and the


tetrahedral A sites in units of Bohr magneton mB . From the experimental values of the Ms, the magnetization by formula unit can be
calculated using the formula [32]:

MMs
5585

100

200

Temperature (K)

300

Fig. 9. Ms dependence on temperature.

(4)

M is the molecular weight.


Fig. 10 exhibits the dependence of the magnetic moment by
formula unit m vs x, calculated using the Eq. (4). A perfect antiferromagnetic alignment for CoFe2O4 in an inverse spinel structure
leads to m 3mB . The normal structure gives m 7mB . S1 presents
m 4:27 mB which is a clear conrmation of the existence of mixed
spinel structure. At 300 K, the value found is fairly agreed with
those reported in the references [17,32]. The difference from literature could be related to the cationic distribution as a result of the
synthesis technique.
Fig. 11 shows the Mr to Msratio, values are around 0.5. For noninteracting particles, Stoner and Wohlfarth expected a remanence
to be half of the saturation magnetization [33]. S1 has a value of 0.6
whereas for S4 and S5 Mr/Ms less than 0.5. For grain sizes smaller
than 20 nm [1], the remanence has been found to be greater of
about 15% relative to 0.5 (20% in this work). This could be related to
the strength of exchange interactions between neighboring grains
[34]. Moreover, S4 and S5 exhibit 0.43 and 0.46 as Mr to Msratio
comparable to Tung and al values [35] obtained for CoFe2O4. As it
was seen in hysteresis loops how surface moments and interparticle interactions manifest, the decrease in the remanence may

Fig. 10. The magnetic moment by formula unit dependence on the cobalt content x.

become evidence. Indeed, when the applied eld shuts to zero


(from its maximum value expecting to saturate the whole moments), the maintaining of the spontaneous magnetization in the
system is ensured by the exchange interactions between neighboring moments rst, second between grains in the case of highly

0,60

80

M /M -10K
R S

MS-10K [emu/g]

100

60

0,52

0,48

0,44

1,0

1,2

1,4

1,6

Cobalt content x
Fig. 8. Ms vs the cobalt content x at 10 K.

1,8

1,0

1,2

1,4

1,6

Cobalt content x
Fig. 11. Mr to Msratio vs the cobalt content.

1,8

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

7. Magnetic anisotropy
According to Mr to Msratio values, it is permitted to assume that
samples have an axial anisotropy [35-37]. For cubic anisotropy, this
ratio is about 0.85 too high than our results. Hence, for a polycrystalline assembly of nanoparticles when neighboring magnetic
exchange interactions can be neglected (Mr/Ms y 0.5) the effective
anisotropy constant K can be deduced from the approach to saturation law and it is given by Ref. [31]:

Keff

1=2

105
b
MS
8

(5)

Here b and Ms have the same signicance as in Eq. 2. Using this


last equation, the effective anisotropy constant Keff has been
deduced for the whole samples and for each temperature. Fig. 12
presents Keff as a function of the temperature for the whole samples. It is clear that Keff is strongly dependent on the temperature
and the cobalt content x. This behavior has been also found by
Adolfo et al [38] for nanoparticles prepared by combustion reaction
method. The difference could be related to the synthesis method
that has been led to different grain sizes which is an important
parameter governing the magnetic properties for single domain
particle. For all samples, the highest anisotropy constant has been
observed at low temperature then it is reduced to its half value at
300 K. In the work of Adolfo et al [38], a much more sensitivity to
temperature has been found,: for example in CoFe2O4, Keff decreases from ~40*106 erg/cm3 at 4 K to ~8*106 erg/cm3 at 272 K for a
grain size about 48 nm, higher than in this work (~7 nm for S1). It
was reported in the literature [39] that CoFe2O4 in its bulk state
exhibits an anisotropy constant being in the range of
2.1e3.9 * 106 erg/cm3 at room temperature. Nevertheless, an
enhancement of Keff has been observed (2e3 times larger) and was
discussed as the resulting from exchange interactions between
neighboring grains [40]. In this work, S1 shows much more
enhancement of K (1.027*107 erg/cm3) at room temperature relative to the bulk value and it is in the order of 10 times larger than
values reported on [38,39]. Hence, it becomes evidence that

x= 1
x=1,2
x=1,4
x=1,6
x=1,8

2,4x10

K (erg/cm3)

2,0x10

1,6x10

1,2x10

reduction grain size would affect inter-particle interactions and


manifests on the macroscopic magnetic properties as the anisotropy energy and the coercive eld.
Unlike in Ref. [38], where the highest anisotropy constant is
found for CoFe2O4, our data grants the highest value for Co1.2Fe1.8O4
when temperature changes from 10 to 300 K. This can be explained
as follow: rst of all, it may advisable to know that the magnetocristalline anisotropy is affected by two competing factors: the
spineorbit coupling and the distortion of the octahedral symmetry
produced by oxygen anions [31]. It has been reported by
Refs. [41,42] that Co2 located on octahedral sites might ensure the
dominant contribution to the anisotropy energy. So, one can believe
that S2 presents the highest Co2 amount which has been lead to
the highest anisotropy constant. Moreover, when replacing Fe3 by
Co2 one affects the spineorbit coupling strength according to the
zero orbital moment of Fe3 [43]. As x increases, the anisotropy
constant drops after reaching its maximum value for S2 (x 1.2). It
is worthy to note that the decrease of the lattice parameter has
been reported to be due to the presence of Co3, the possible
migration of Co2 from B to A sites and the possible existence of
vacancies. These causes will have a remarkable effect on the crystal
eld (shrinkage of the lattice parameter leads to a distortion of the
octahedral crystal eld). Then a decrease of the strength of the spin
orbit coupling, originates from the diamagnetic behavior of Co3
0 mB [31], takes place.
For monodomainenanoparticles, the magnetocristalline anisotropy constant can be deduced from ZFC measurements using
TBvalues and the Neel-Brown relaxation law written as follow:

KZFC

25KB TB
V

(6)

where KB is the Boltzmann constant and V is is the mean volume


[44]. Fig. 13 shows KZFC vs x. When increasing x, KZFC shows a
weakly variation till x 1.4 and a rapid increase for x 1.6 and 1.8.
This dependence has been found in our group by Ajroudi et al [45].
One comparing the kinds of estimation, the second one is much
more sensitive to the size of the system. This dependence of KZFC on
x is a clear evidence of nite size effects. Probably, using the law of
approach to saturation the additional anisotropies as shape, strain
and surface anisotropies were underestimated. For the two kinds of
anisotropies, values are close but with two different behaviors.
At this stage, one can predict that the behavior found on the
hysteresis shape for x 1.6 and 1.8 is due to the existence of surface
effects.

1,4x10

KEFF erg/cm3

interacting particles. So, surface moments don't have the sufcient


exchange interaction resulting from the lack of neighboring moments resulting of a rapidly decrease relative to the core. That is
why S4 and S5 present the smallest remanence.

519

7,0x10

8,0x10

40

80

120

160

200

Temperature (K)

240

280

Fig. 12. effective anisotropy constant for CoxFe3-xO4 samples as a function of the
temperature.

1,0

1,2

1,4

1,6

cobalt content x
Fig. 13. Effective anisotropy deduced from ZFC measurements.

1,8

520

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

8. Surface effects
For ordered magnetic systems, the temperature dependence of
the magnetization is related to the presence of low energy collective excitations known as magnons [46]. This phenomenon manifests by a decrease of the spontaneous magnetization as the
temperature increases. Thus, it allowed the development of a power law that describes the dependence of the saturation magnetization on the temperature known as Bloch T 3=2 law and given by
Refs. [35,47]:

MS T MS 01  BT a

(7)

where MS 0 is the saturation magnetization estimated at 0 K and B


and a are the Bloch constant and exponent which are size and
structure dependent [48,49]. It was reported in literature that B
increases as the size decreases and Bsurface may be in order of
magnitude 10 times larger than BBulk . For bulk CoFe2O4, NiFe2O4
and Fe3O4 [2] a is equal to 2 and ts well the experimental data.
Nevertheless, a value of 1.5 has been found to also t Ms (T) in
MgxFe3xO4 [48] for sizes range from 35 to 50 nm which is higher
than in this work. Hence, it is clear that Bloch's law parameters are
strongly structure and atomic features dependent as well as the
exchange energy [9] which has been extracted from Bloch constant
and found to be dependent on the Mg content [48].
Fig. 9 displays the tting of the Ms using the Eq. (5). Fig. 14
shows the dependence of a and B on x: as x increases, a decreases. The highest value, 1.75, is assigned to S1 (CoFe2O4) and the
lowest value, 1.39 to S5 (Co1.8Fe1.2O4). B has been found to increase
with increasing x. S1 has a value of 9.97 106 K1.75, S5 has a value
of 9 105 K1.39. So, a decreases and BeS5 is about 10 times larger
than BeS1. From these results, we believe that the observed jump
in S4 and S5 is mainly due to surface effects and are more pronounced in S5. The dependence of a on the particle's diameter has
been studied by Ngo et al [50] and the results are similar to our
results; an increase of a with increasing particle size.
9. Coercive eld and interparticle interactions

T
HC HC0  1 
TBK

0:5 !
(8)

HC0 and TBK are the coercive eld at 0 K and the superparamagnetic
blocking temperature. From this t, the Kneller's blocking temperature (TBK) has been extracted and the results are plotted in
Fig. 17. One can note that the results are in fairly agreement with
values estimated using ZFC for all cobalt content except for x 1.6
where Kneller's law t value deviates from the value estimated
form ZFC value. This discrepancy reects the strength of the
interparticle interaction [21]. Because S1-ZFC's curve presents the
sharper peak, the difference between the value estimated from ZFC
curve and the value extracted from the Kneller's law t is the

12

HC (kOe)

From M (H) curves, the coercive eld HC has been extracted and
the data are plotted in Fig. 15 (symbols). It can be shown that HC
depends on both the temperature and the cobalt content. Such
behavior has been observed for CoFe2O4 nanoparticles [2] and for
manganese-substituted cobalt ferrite by Melikhov et al [42]. HC
increases when the temperature decreases for all the samples
except S4 (1.6) which exhibits a weak value relative to the other

samples even at 10 K. This out of range value is due to the presence


of aggregated nanoparticles as shown by TEM image (Fig. 3c).
The main parameters inuencing the coercive eld are the grain
size [51] and the microstructural features, although, the magnetic
exchange interactions [52] which depend on the cationic distribution. Sample's sizes are found to be weakly dependent on the
cobalt amount whereas the dependence of the coercive eld on the
size seems to be not totally affected at 10 K. Samples S1, S2 and S3
exhibit very similar values 10.7, 9.5 and 11 kOe respectively, at 10 K
and found to be higher than those reported by Refs. [2,21]. Referring to M (H) curves recorded at 300 K (Fig. 7d), S2, S3, S4 and S5
have no hysteresis. This feature is the footprint of the superparamagnetism [51] behavior at room temperature. The dependence of the coercive eld on the grain size will be governed by the
volume and the anisotropy energy as it was described by Neel [51].
Fig. 16 shows HC dependence of the cobalt content x at 300 K. It is
clear that the cobalt content has affected the coercive eld through
the anisotropy energy unlike the behavior found for 10 K.
Fig. 17 presents the blocking temperature (TB) as a function of
the cobalt content x. It is estimated as the maximum of ZFC curves
for single non-interacting mono-domain nanoparticles [2]. TB
almost decreases with x originate from the overcoming of the
barrier anisotropy energy KV by the thermal uctuations. The
decrease of TB is mainly related to the decrease of the anisotropy
effective constant K.
For non-interacting single domain magnetic nanoparticles, the
dependence of the coercive eld on the temperature has been
found to follow an empirical law known as the Kneller's law [52]
given by:

x=1
fit x=1
x=1,2
fit x=1,2
x=1,4
fit x=1,4
x=1,6
fit x=1,6
x=1,8
fit x=1,8

100

200

Temperature (K)
Fig. 14. Bloch's law constants dependence on x.

Fig. 15. Coercive eld dependence on the temperature for CoxFe3XO4.

300

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

H -300 K (kOe)
C

0,3

521

diamagnetic cations which act on the exchange energy CoeOeFe.


An agglomerated particles like have been observed for S4 (x 1.6)
by TEM. The magnetic response has been conrmed by its effect on
the coercive eld, the aggregated nanoparticles have allowed to
high interparticle interactions. The interparticle interactions have
been conrmed by means of the discrepancy found between the
Kneller blocking temperature, estimated from the Kneller's law t
of the coercive eld dependence on the temperature, and the
blocking temperature TB.
In the future, we plan to study the surface effects, the cationic
distribution by means of the 57Fe Mossbauer Spectroscopy.

0,2

0,1

Acknowledgment

0,0
1,0

1,2

1,4

1,6

1,8

Cobalt content x
Fig. 16. HC dependence on the cobalt content x at 300 K.

TB from ZFC measurment

Temperature (K)

320

TB from Kneller's law fit

300

re de l'Enseignement
This work was supported by the Ministe
rieur, de la Recherche Scientique, Tunisia and l'Agence
Supe
Universitaire de la Francophonie (AUF) via the scholarship program
Eugen Ionesco. We thank also the Sectoral Operational Programme Human Resources Development (SOP HRD), nanced from
the European Social Fund and by the Romanian Government under
the project number POSDRU/159/1.5/S/134378 and by the structural founds project PRO-DD (POS-CCE, O.2.2.1., ID 123, SMIS 2637,
ctr. No 11/2009) for providing the infrastructure used in this work.
References

280
260
240

1,0

1,2

1,4

1,6

1,8

Cobalt content x
Fig. 17. Blocking temperature versus cobalt content x for CoxFe3xO4 samples.

minimum. One notices the existence of a non-neglected disparity


for x 1.2 (S2). 10 K hysteresis loop for S2 displays a drop nearby
0 applied led (see Fig. 7E). This result conrms the discrepancy
between TB and TBK and it originates from interparticle interactions
as discussed above. S1, S3 and S5 display weak deviation relative to
S2 and S4 reecting neglected or very weakly interparticle
interactions.
10. Conclusions
We have studied the effect induced by the substitution of cobalt
by iron following the formula CoxFe3xO4 (1  x< 1.8) using a solvothermal route. We have found that the higher cobalt content
(x 1.6 and 1.8) samples (S4 and S5) is the subject of nite sizes and
surface effects. At low temperature, the moments at the surface acts
on the magnetization by a jump on the hysteresis loops, a rapid
decrease of the magnetization nearby zero applied elds. We have
conrmed the surface effects by the magnetic anisotropy estimated
from ZFC measurements and the tting of the saturation magnetization dependence on the temperature to the power Bloch T3/2
law. B and a, which are the Bloch's law constants, have been found
to be very dependent on x. The blocking temperature and the
magnetocristalline anisotropy energy have been found to decrease
with increasing x. This has been related to the presence of Co3 as

[1] A. Manaf, R.A. Buckley, H.A. Davies, M. Leonowicz, J. Magn. Magn. Mater. 360
(101) (1991), http://dx.doi.org/10.1016/0304e8853(91)90779-A.
[2] C. Vazquez-Vazquez, M.A. Lopez-Quintela, M.C. Bujan-Nunez, J. Rivas,
J. Nanopart. Res. 13 (2011) 1663, http://dx.doi.org/10.1007/s11051-010-99207.
[3] N. Sanpo, C.C. Berndt, C. Wen, James Wang, Acta Biomater. 9 (2013) 5830,
http://dx.doi.org/10.1016/j.actbio.2012.10.037.
[4] Q.A. Pankhurst, J. Connolly, S.K. Jones, J. Dobson, J. Phys. D. Appl. Phys. 36
(2003) R167, http://dx.doi.org/10.1088/0022e3727/36/13/201.
[5] A. Franco Jr., V.S. Zapf, V.B. Barbeta, R.F. Jardim, J. Appl. Phys. 107 (2010)
073904. http://dx.doi.org/10.1063/1.3359709.
[6] A.K. Giri, E.M. Kirkpatrick, P. Moongkhamklang, S.A. Majetich, V.G. Harris Appl,
Phys. Lett. 80 (2002) 2341, http://dx.doi.org/10.1063/1.1464661.
[7] C.N. Chinnasamy, M. Senoue, B. Jeyadevam, O. Perales-Perez, K. Shinoda,
K. Tohji, J. Colloid Interface Sci. 263 (2003) 80, http://dx.doi.org/10.1016/
S0021-9797(03)00258-3.
[8] J. Stohr, H.C. Siegmann, Magnetism from Fundamentals to Nanoscale Dynamics, springer, 2006.
[9] G.A. Sawatzky, F. van der Woude, A.H. Morrish, J. App. Phys. 39 (1968) 1204,
http://dx.doi.org/10.1063/1.1656224.
[10] J.A. Moyer, C.A.F. Vaz, E. Negusse, D.A. Arena, V.E. Henrich, Phys. Rev. B 83
(2011) 035121, http://dx.doi.org/10.1103/PhysRevB.83.035121.
[11] G. Hu, J.H. Choi, C.B. Eom, V.G. Harris, Y. Suzuki, Phys. Rev. B 62 (2000) R779,
http://dx.doi.org/10.1103/PhysRevB.62.R779.
[12] M. Mozaffari, J. Amighian, E. Darsheshdar, J. Magn. Magn. Mater. 350 (2014)
19, http://dx.doi.org/10.1016/j.jmmm.2013.08.008.
[13] L. Ajroudi, S. Villain, V. Madigou, N. Mliki, Ch. Leroux, J. Cryst. Growth 312
(2010) 2465e2471, http://dx.doi.org/10.1016/j.jcrysgro.2010.05.024.
[14] L. Vegard, Z. Phys. 5 (1921) 17.
[15] S.K. Pradhan, S. Bidb, M. Gateshki, V. Petkov, Mater. Chem. Phys. 93 (2005)
224, http://dx.doi.org/10.1016/j.matchemphys.2005.03.017.
[16] Dong Hoon Lee, Hong Seok Kim, Jeong Yong Lee, Chul Hyun Yo, Keu
Hong Kim, Solid State Commun. 96 (1995) 445, http://dx.doi.org/10.1016/
0038-1098(95)00491-2.
[17] I. Sharif, H. Shokrollahi, M.M. Doroodmand, R. Sa, J. Magn. Magn. Mater. 324
(2012) 1854, http://dx.doi.org/10.1016/j.jmmm.2012.01.015.
[18] H. LeTrong, L. Presmanes, E. .DeGrave, A. Barnabe, C. Bonningue, Ph. Tailhades,
J. Magn. Magn. Mater. 334 (2013) 66, http://dx.doi.org/10.1016/
j.jmmm.2013.01.007.
[19] F. Nakagomi, S.W. da Silva, V.K. Garg, A.C. Oliveira, P.C. Morais, A. Franco Jr.,
J. Solid State Chem. 182 (2009) 2423, http://dx.doi.org/10.1016/
j.jssc.2009.06.036.
[20] Ch. C. Dantas, A.M. Gama, J. Magn. Magn. Mater. 322 (2010) 2824, http://
dx.doi.org/10.1016/j.jmmm.2010.04.037.
[21] Y. Zhang, Y. Liu, C. Fei, Z. Yang, Z. Lu, R. Xiong, D. Yin, J. Shi, J. App. Phys. 108
(2010) 084312, http://dx.doi.org/10.1063/1.3499289.
[22] A. Raghunathan, P. Klimczyk, Y. Melikhov, IEEE Trans. Magnet. 49 (2013) 7,
http://dx.doi.org/10.1109/TMAG.2013.2243823.
[23] A.B. Antunes, M.N. Baibich, V. Gil, C. Moure, V. Allegret-Maret, O. Pena,
J. Magn. Magn. Mater 320 (2008) e464, http://dx.doi.org/10.1016/
j.jmmm.2008.02.085.

522

M. Saidani et al. / Journal of Alloys and Compounds 653 (2015) 513e522

[24] R.D. Zysler, C.A. Ramos, E. De Biasi, H. Romero, A. Ortega, D. Fiorani, J. Magn.
Magn. Mater. 221 (2000) 37, http://dx.doi.org/10.1016/S0304-8853(00)
00368-1.
[25] Dino Fiorani, Surface Effects in Magnetic Nanoparticles, page 92
[26] C.R. Alves, R. Aquino, J. Depeyrot, T.A. P Cotta, M.H. Sousa, F.A. Tourinho,
H.R. Rechenberg, G.F. Goya, J. Appl. Phys. 99 (2006) 08M905. http://dx.doi.org/
10.1063/1.2163844.
[27] C. Cannas, A. Falqui, A. Musinu, D. Peddis, G. Piccaluga, J. Nanopart. Res. 8
(2006) 255, http://dx.doi.org/10.1007/s11051-005-9028-7.
[28] A. Franco Jr., V.S. Zapf, V.B. Barbeta, R.F. Jardim, J. Appl. Phys. 107 (2010)
073904, http://dx.doi.org/10.1063/1.3359709.
[29] T. Gaudisson, M. Artus, U. Acevedo, F. Herbst, S. Nowak, R. Valenzuel,
S. Ammar, J. Magn. Magn. Mater. 370 (2014) 87, http://dx.doi.org/10.1016/
j.jmmm.2014.06.014.
[30] F. Nakagomi, S.W. da Silva, V.K. Garg, A.C. Oliveira, P.C. Morais, A. Franco Jnior, E.C.D. Lima, J. Appl. Phys. 101 (2007) 09M514, http://dx.doi.org/10.1063/
1.2712821.
[31] I.C. Nlebedim, A.J. Moses, D.C. Jiles, J. Magn. Magn. Mater. 343 (2013) 49,
http://dx.doi.org/10.1016/j.jmmm.2013.04.063.
[32] R.C. Kambale, P.A. Shaikh, S.S. Kamble, Y.D. Kolekar, J. Alloys Comp. 478 (2009)
599, http://dx.doi.org/10.1016/j.jallcom.2008.11.101.
[33] E.C. Stoner, E.P. Wohlfarth, Phil. Trans. Roy. Soc. 240 (1948) 599, http://
dx.doi.org/10.1098/rsta.1948.0007.
[34] R. Fisher, T. Shre, H. Kronmuller, J. Fidler, J. Magn. Magn. Mater 153 (1996)
35, http://dx.doi.org/10.1016/0304-8853(95)00494-7.
[35] L.D. Tung, V. Kolesnichenko, D. Caruntu, N.H. Chou, C.J. O'Connor, L. Spinu,
J. Appl. Phys. 93 (2003) 7486, http://dx.doi.org/10.1063/1.1540145.
[36] D.H. Manh, P.T. Phong, T.D. Thanh, D.N.H. Nam, L.V. Hong, N.X. Phuc, J. Alloys
Compd. 509 (2011) 1373, http://dx.doi.org/10.1016/j.jallcom.2010.10.104.
[37] R. Fischer, H. Kronmuller, Phys. Rev. B 54 (1996) 7284. http://dx.doi.org/10.

1103/PhysRevB.54.7284.
[38] Adolfo Franco Jr., V. Zapf, J. Magn. Magn. Mater. 320 (2008) 709, http://
dx.doi.org/10.1016/j.jmmm.2007.08.009.
[39] H. Shenker, Phys. Rev. 107 (1957) 1246, http://dx.doi.org/10.1103/
PhysRev.107.1246.
[40] Kim E. Mooney, J.A. Nelson, M.J. Wagner, Chem. Mater. 16 (2004) 3155, http://
dx.doi.org/10.1021/cm040012.
[41] M. Tachiki, vol. 23, no.6, June 1960, 10.1143/PTP.17.331.
[42] Y. Melikhov, J.E. Snyder, D.C. Jiles, A.P. Ring, J.A. Paulsen, et al., J. Appl. Phys. 99
(2006) 08R102, http://dx.doi.org/10.1063/1.2151793.
[43] N. Moumen, P. Bonville, M.P. Pileni, J. Phys. Chem. 100 (1996) 14410, http://
dx.doi.org/10.1021/jp9524136.
el, The
orie du tranage magne
tique des ferromagne
tiques en grains ns
[44] L. Ne
ophys. 5 (1949) 99.
avec applications aux terres cuites, Ann. Ge
[45] L. Ajroudi, N. Mliki, L. Bessais, V. Madigou, S. Villain, Ch. Leroux, Mater. Res.
Bull. 59 (2014) 49e58, http://dx.doi.org/10.1016/j.materresbull.2014.06.029.
[46] S. Cojocaru, Solid State Commun. 151 (2011) 1780, http://dx.doi.org/10.1016/
j.ssc.2011.08.028.
[47] D. Zhang, K.J. Klabunde, C.M. Sorensen, G.C. Hadjipanayis, Phys. Rev. B 58
(1998) 14167. http://dx.doi.org/10.1103/PhysRevB.58.14167.
[48] A. Franco Jr., V.S. Zapf, V.B. Barbeta, R.F. Jardim, J. Appl. Phys. 107 (2010)
073904. http://dx.doi.org/10.1063/1.3359709.
[49] R. Aquino, J. Depeyrot, M.H. Sousa, F.A. Tourinho, E. Dubois, R. Perzynski, Phys.
Rev. B 72 (2005) 184435. http://dx.doi.org/10.1103/PhysRevB.72.184435.
[50] A.T. Ngo, P. Bonville, M.P. Pileni, J. Appl. Phys. 89 (2001) 3370. http://dx.doi.
org/10.1063/1.1347001.
[51] C.P. Bean, J.D. Livingston, J. Appl. Phys. 30 (1959) S120. http://dx.doi.org/10.
1063/1.2185850.
[52] E.F. Kneller, F.E. Luborsky, J. Appl. Phys. 34 (1963) 656. http://dx.doi.org/10.
1063/1.1729324.

You might also like