You are on page 1of 13

Applied Physics A (2020) 126:660

https://doi.org/10.1007/s00339-020-03823-9

Surfactant based synthesis and magnetic studies of cobalt ferrite


Meenal Gupta1   · Anusree Das2   · Satyabrata Mohapatra1   · Dipankar Das2   · Anindya Datta1 

Received: 27 March 2020 / Accepted: 18 July 2020


© Springer-Verlag GmbH Germany, part of Springer Nature 2020

Abstract 
Nanoparticles of cobalt ferrite were synthesized by two different surfactants cetyl trimethyl ammonium bromide (CTAB)
and sodium dodecyl sulphate (SDS) by the microemulsion method. Formation of cobalt ferrite was confirmed by structural
characterization techniques XRD and TEM. The magnetic properties of these systems were studied using Mössbauer spec-
troscopy and SQUID magnetometry. The study determined the impact of the synthetic procedure and surfactant composi-
tion on the nanomaterial characteristics. The overall magnetic response for the material was different compared to the bulk
with a major decrease in room temperature magnetic saturation. Magnetic properties were affected by the type of surfactant
template due to its impact on the geometry of the synthesized cobalt ferrite nanoparticles.

* Anindya Datta
anindya_datta@yahoo.com
Meenal Gupta
meenal_garg04@hotmail.com
Anusree Das
anusreedas8@gmail.com
Satyabrata Mohapatra
smiuac@gmail.com
Dipankar Das
ddas.ugcdae@gmail.com
1
BFR-206, B Block, University School of Basic and Applied
Sciences, Guru Gobind Singh Indraprastha University,
Dwarka, Delhi, India 110078
2
UGC-DAE CSR, Bidhan Nagar, Kolkata, West Bengal, India

13
Vol.:(0123456789)
660   Page 2 of 13 M. Gupta et al.

Graphic abstract

13
Surfactant based synthesis and magnetic studies of cobalt ferrite Page 3 of 13  660

Keywords  Cobalt ferrite · Surfactant · Magnetic · Mössbauer spectroscopy

1 Introduction with the main aim to fine-tune the properties like particle
size and distribution, shape, degree of agglomeration and
Magnetic properties associated with materials are heavily particle composition. Several methods such as sol–gel,
modified as these approach nanoscale. Large surface area co-precipitation, microemulsion, etc. have been reported
and quantum nanosize effect contribute towards superpara- for ferrite synthesis [14–19]. Aubery et al. [20] prepared
magnetism and quantum tunnelling [1]. Ferrite magnetic Mn − Zn ferrite using the microemulsion method using Syn-
nanoparticles have been found to be crucial materials for peronic13 surfactant and isooctane oil medium. They stud-
application in magnetic resonance imaging (MRI), ferrofluid ied reaction kinetics using small-angle neutron scattering
technology, guided drug delivery, magnetocaloric refrigera- (SANS) and dynamic light scattering (DLS) and synthesized
tion among others. Among ferrites, the spinel ferrites are the nanoparticles of sizes 4 nm and 8 nm by varying reaction
source of immense interest because of their cubic symmetry conditions. Moumen and Pileni reported the formation of
[2–4]. Spinel structure was first observed by Bragg in natu- nanosized cobalt ferrite by the micelle method using SDS
ral form in ­MgAl2O4[5]. This type of lattice arrangement as surfactant with 2–5 nm as the average particle size that
was observed and further studied for magnetic nanoparticles could be controlled by varying the surfactant concentration
such as various ferric oxides or ferrites [6, 7]. The natu- [21]. Seip et al. [22] also used the same methodology but
ral preference of oxides for the spinel FCC structure due to reverse micelles to synthesise different ferrites using AOT
its stability has inspired the study of the impact on various instead of SDS and studied their magnetic properties. In both
chemical and physical properties of spinel ferrites by intro- cases, the particles exhibited superparamagnetic character in
ducing disturbance to such an ordered system [8]. Cobalt 2–5 nm size range. Maaz et al. [23] used oleic acid as a sur-
ferrite nanoparticle has a mixed inverse spinel structure with factant to prepare nanosized cobalt ferrite by the wet chemi-
divalent cobalt cations and trivalent ferric cations distributed cal route. The particle size range was 15–20 nm as deduced
at tetrahedral (A) and octahedral (B) sites in its lattice [9]. from XRD and TEM. In microemulsion based chemical
Its chemical stability and magnetic hardness facilitate its synthetic technique, uniform and controlled composition
applications in magnetic refrigeration, high-density record- and morphology of nanoparticles are achieved because of
ing, spintronics, ferrofluids, anti-tumour drug delivery and reagent cation confinement and particle stabilization by
many others. Cobalt ferrites exhibit large magnetocrystalline surfactant molecules [24]. The surfactant molecules created
anisotropy making them important technological materials soft template matrix contains several nanoreactors enclosed
and they also have considerable magnetization [10, 11]. within. The surfactant molecules because of their shape,
The spinel structure is of three types: normal, inverse charge distribution, head and tail groups create a different
and mixed based on cationic distributions of the divalent type of micro-reactors. The nanoparticle formation occurs
­M2+ and trivalent ­Fe3+ in octahedral and tetrahedral sites in inside these nanoreactors and hence the size and geometry
the cubic structure. The ferrimagnetic M ­ Fe2O4 (M = Co, Ni) of these reactors influence the structural properties of the
nanoparticles have spinel cubic structure with two sublat- nanoparticles. These surfactant imprints preserve even after
tices called A for tetrahedral site and B for octahedral site the surfactant is eliminated from the system [25]. The inter-
and belong to the Fd3m space group. Cobalt ferrite is an action as a result of the attractive or repulsive force between
inverse spinel with general formula F ­ e3+A(Co2+ ­Fe3+)BO4 the head and the nanoparticle due to the polarities of the
where trivalent cation is equally distributed at octahedral and particle and the surfactant head have detrimental effect on
tetrahedral sites and divalent cation occupies the octahedral the nanoparticle growth [26, 27].
site [12, 13]. In the present work, we have synthesized cobalt ferrite
Synthesis methods for ferrite nanoparticles include both nanoparticles using reverse micelle nanoreactors in the
top-down and bottom-up approaches. In top-down, the fun- water-in-oil microemulsion system using two different sur-
damental process is the conversion of the bulk material factants of similar stoichiometry but different polarities.
into smaller pieces in the presence of external energy that The cationic metal ion interacts differently with the anionic
may be mechanical or chemical in nature. Using atomic or (SDS) and cationic (CTAB) surfactant causing different
molecular precursors, bottom-up approach utilises chemical kinds of nano aggregates in both cases. With equivalent
reactions to synthesize and nanoparticles and allows the pre- hydrophobic tails, the primary effect on the nanoparticle
cursor particles to grow. Both methodologies can be done in formation is governed by the micellization of the head
either solid, gas, liquid, supercritical fluids, or in a vacuum around it. It was reported that similarly charged components
experience electrostatic repulsion while oppositely charged

13
660   Page 4 of 13 M. Gupta et al.

2 Materials and methods

Wet chemical synthesis of cobalt ferrite nanoparticles by


the use of surfactants was carried out. Precursor salts of
Iron (III) and Cobalt (II) were mixed in ratio of 2:1, respec-
tively, in deionised water and gradually added to 50 mL of
Surfactant /n-Hexane/Butanol solution in stoichiometric
amounts to form a continuous phase. The reaction mixture
was continuously kept stirring for 20 h. After that, sodium
hydroxide was added dropwise to the microemulsion to
bring the pH in the range of 10. The mixture was heated at
80 °C till slurry formation was achieved with the addition
of 7 mL of triethylamine. After washing with ethanol and
water, the mixture was dried overnight in an oven at 80 °C
and then calcined at 550 °C for three hours.
The samples synthesized by SDS/n-Hexane/Butanol were
named A1, A2 and A3 and CTAB/n-Hexane/Butanol were
named B1, B2 and B3 in order of increasing surfactant con-
Fig. 1  XRD patterns for cobalt ferrite prepared using different SDS
concentrations
centrations of 0.08 M, 0.10 M and 0.14 M.
The samples were characterized using XRD, TEM,
SQUID and Mössbauer spectroscopy. The XRD patterns
were recorded by a D8 Advance diffractometer (Bruker,
Germany) using a Cu K-alpha source with 1.54 Å wave-
length. TEM images were captured on JEOL 2100 at CRNN,
University of Calcutta. SQUID measurements were per-
formed using Quantum Design, USA make MMPS XL 7
magnetometer and Mössbauer spectroscopy was studied at
room temperature using a 10 mCi Co-57 in Rh matrix as the
gamma-ray source.

3 Results and discussion

XRD and TEM techniques are utilised for structural analysis


of the samples.
Figures 1 and 2 show the XRD patterns for cobalt ferrite
synthesized using three different surfactant concentrations
of 0.08 M, 0.10 M and 0.14 M of SDS and CTAB, respec-
tively. The patterns are in agreement with JCPDS #22–1086
Fig. 2  XRD patterns for cobalt ferrite prepared using different CTAB which corresponds to crystalline inverse spinel cobalt ferrite.
concentrations
The most intense peak for (311) is observed at 35.56O. The
value of lattice constant is found to be slightly lower than the
experience attractive forces. This impacts assemblage of reported value at 8.3 Å [23]. The crystallite size is calculated
surfactant micelles and thus the microreactors associated from the most intense (311) peak using the Debye–Scherrer
with the microemulsion system [28].Our study is focussed equation [29]:
on the impact on the magnetic response of the cobalt ferrite
nanoparticles as a consequence of the two different micellar 09𝜆
Cs = (1)
systems used for synthesis and acting as templates and the 𝛽 cos θ
preservation of effects of surfactant on the lattice above the where Cs is the crystallite size and λ is the Cu wavelength
vaporisation temperatures of the chosen surfactants. i.e. 1.54 Å, β is the full width at half maximum and θ is the
Bragg angle.

13
Surfactant based synthesis and magnetic studies of cobalt ferrite Page 5 of 13  660

Sch 1  Use of CTAB and SDS in Surfactant/n-Hexane/Butanol Microemulsion for Cobalt 496 Ferrite Nanoparticle Synthesis

Table 1  XRD results for samples A1-A3 and B1-B3


Sample 2Ɵ (degree) Lattice con- Crystallite X-ray density,
stant, a(Å) size (nm) ­(dx) (g ­cm−3)

A1 35.8 8.309 10.5 5.4


A2 35.8 8.310 14.1 5.4
A3 35.8 8.309 13.1 5.4
B1 35.9 8.316 13.5 5.5
B2 35.7 8.326 14.3 5.5
B3 35.7 8.336 13.6 5.4

The ferrite nanoparticles are formed by the interaction


of ferric and cobaltous salts in the nanopools that are sus-
pended in the oil phase. The interaction between the two
cations occurs in aqueous phase enclosed in these nano-
pools or nanoreactors mentioned above. The reduction of
surface tension is the main purpose of the surfactant at the Fig. 3  Pre-heat treatment XRD pattern for as synthesized cobalt fer-
interface between oil and water layers (that are the walls of rite
the nanoreactors in this case) to enable the formation of the
mixed metal hydroxide compound. The type of surfactant
determines the geometry of the nanoreactor/ nanopool which concentration (CMC) values as below CMC no micelle
influences the morphology and size of the nanoparticles formation and thus no microemulsion system is formed.
formed as described in Scheme 1 [30]. For higher surfactant Also higher surfactant amount ensures better kinetics as the
concentration, the samples tend to be more crystalline as amount of secondary and amorphous phases is observed to
evident from the XRD patterns. Increasing the surfactant be considerably lower when going from samples 1–3 for
concentration introduces more nanoreactors into the system both A (SDS) and B (CTAB) series [32]. The difference in
leading to the formation of larger micellar assembly. When particle sizes among SDS and CTAB samples is because
the size of nanoreactor increases, the particle size is also of different nano assemblies for each. The CMC value of
varied with it as seen in Table 1 [31]. The concentrations CTAB is lower than that of SDS leading to early micelliza-
of surfactants are above their respective critical micellar tion in same molarity CTAB microemulsion. The time for

13
660   Page 6 of 13 M. Gupta et al.

crystalline particles are in nanoscale. The average d-spacing


for samples A3 and B3 is about 2.5 Å which is the reported
value for (311) plane that is represented by the brightest
ring in the SAED pattern. From these TEM micrographs,
the sizes of nanoparticles of the two samples are measured
and histograms plotted that are found to be best fitted with
the lognormal distribution [41]. From these histogram plots
in Fig. 7, the average nanoparticle diameters calculated for
samples A3 and B3 are 12.2 ± 3.4 nm and 13.1 ± 4.1 nm,
respectively.
With the change in the surfactant amount, the size of
nanoparticle is affected as the size of nanopool is depend-
ent on the ratios of oil/surfactant/water concentrations [42].
The type of surfactant concentrations taken for both types
of micellar systems are the same although the overall mor-
phology is distinct as the particle shapes tend to differ. This
difference arises due to two distinct mechanisms of particle
formation. SDS is an anionic surfactant. Therefore, in SDS
Fig. 4  a–d Pre-heat treatment TEM and SAED pattern for A3(a, b) mediated system, the mechanism of particle formation as
and B3(c, d), respectively
described [9, 43] is due to the formation of aqueous cobalt-
iron disulphide that converts to cobalt iron oxide in basic
particle growth is more and hence B series particles have medium with addition of amine. On the other hand, CTA​+ is
larger sizes. The use of anionic surfactants produces smaller the long-chain part of the cationic CTAB surfactant that gets
sized nanoparticles compared to cationic surfactants [33]. attached to the surface of the cobalt-iron hydroxide which
The length of the hydrophobic chain directly affects the par- inhibits further particle growth to give uniform cobalt ferrite
ticle size [34, 35]. Among the various concentrations of the nanoparticles [44, 45]. From the results of XRD and TEM,
same surfactant, a linear relationship between particle size the average particle sizes for both samples are below the
and surfactant molarity is expected. Increment in surfactant critical size for cobalt ferrite nanoparticles with the exist-
amount leads to an increase in the size of microreactors and ence of multi domains. This leads to interesting results for
hence the size of the nanocluster also increases. Excess sur- magnetic properties compared to the bulk sample and the
factant causes the presence of excess hydrophobic ligands reported results [2, 42, 46].
that inhibit the particle growth [36, 37]. Leff et al. [38] has Figure 8a–d show Mössbauer spectra at room temperature
reported an inverse relation between surfactant concentra- for cobalt ferrite nanoparticles prepared via the microemul-
tion and particle size in the presence of excess surfactant sion route using two different surfactants. The Mössbauer
amounts and has attributed it to positive entropy due to sur- effect is based on the interactions of the nuclear charge
factant adsorption. Particle size reduces to compensate and with the electromagnetic field created by the surrounding
reduce entropy. extranuclear electrons [47]. Being a local probe, it gives
Consider Figs. 3 and 4 containing XRD pattern and TEM information on the surrounding of the nucleus. The three
micrographs for samples A3 and B3 before calcination. experimentally measurable parameters in Mössbauer spec-
Here,  additional phases are seen in the XRD pattern. As a troscopy are isomer shift, quadrupole splitting and hyperfine
result of synthesis temperature being lower than the vapori- (internal) magnetic field. The isomer shift gives information
sation temperature for SDS and CTAB, small quantities of on the valence state of Fe atoms in the sample. The quadru-
surfactant is still present in the basic morphology of the pole splitting gives information on the symmetry of charge
nanostructure. Both types of nanostructures show different distribution around the probe nucleus whereas the internal
nano assemblies due to different aggregation mechanism for magnetic field provides a measure of magnetic ordering in
each type of surfactant micelle [39, 40]. Due to the presence the sample [48]. The obtained Mössbauer spectra for SDS
of surfactants in the system, the average nanoparticle size and CTAB mediated cobalt ferrite nanoparticles are fitted
is about 17.2 nm for SDS samples and 17.7 nm for CTAB using least-squares fits of Lorentzian function to two sextets
samples. To obtain a complete pure sample, this sample was and a doublet. It may be noted that the sextet patterns show
heated at 550 °C in the furnace. The TEM of the calcined a signature of relaxation as the peak intensity ratio differs
samples do not show surfactant impurity. from being 3:2:1: 1:2:3 expected for a powder sample with
From the  TEM micrographs for cobalt ferrite speci- magnetic ordering. In addition, line widths of the peaks are
mens A3 and B3 as seen in Figs. 5 and 6, it is clear that the very large confirming the presence of superparamagnetic

13
Surfactant based synthesis and magnetic studies of cobalt ferrite Page 7 of 13  660

Fig. 5  a TEM image, b–c HRTEM images and d SAED pattern for sample A3

relaxation in the samples. From Table 2, isomer shift values vary with particle size, decreasing with a decrease in particle
are in the range of 0.31 to 0.34 mm s−1 for the sextets for size. The presence of a central doublet confirms superpara-
samples A2 and A3 and 0.26 to 0.37 mm s−1 and for B2 and magnetic behaviour. This impact of change in surfactant is
B3, respectively. The isomeric shift values obtained for both witnessed in isomeric shift and hyperfine field values due to
are lower in comparison to the reported values at 300 K [10]. surface morphology as well as crystal quality as a result of
This is due to the superparamagnetic behaviour of cobalt fer- the synthetic procedure [12, 51].
rite nanoparticles that is verified by the absence of hysteresis From the results of magnetic measurements conducted
curve for these nanoparticles at the same temperature using by SQUID magnetometry at room temperature and 5 K as
SQUID that will be discussed in the later section. Reduced shown in Fig. 9, it is clear that prominent hysteresis curves
isomeric shift values for A2-A3 and B2-B3 with particle are obtained at 5 K while no hysteresis is seen at 300 K. The
sizes between 10–20 nm compared to bulk cobalt ferrite size of cobalt ferrite nanoparticles is less than the super-
arises due to increase in density of s-electrons of the Fe atom paramagnetic diameter ­(dSP) for cobalt ferrite [2, 52] which
because of decrease in distance between ­Fe3+–O2− ligands is also seen in the room temperature hysteresis. Due to the
at octahedral and tetrahedral sites with a decrease in size small size, the magnetic nanoparticles cannot accommodate
[49, 50]. The increase in surface area with a decrease in size multi-domain system which reduces the corresponding mag-
leads to distortion of the crystal lattice that affects the elec- netostatic energy. The magnetostatic energy or the activa-
tron density and spin orientation. The hyperfine field values tion energy is thus less than the thermal energy at 300 K.

13
660   Page 8 of 13 M. Gupta et al.

Fig. 6  a TEM image, b–c HRTEM images and d SAED pattern for sample B3

According to the theory of superparamagnetism (SPM), all to destroy such alignment [56]. A close look at Fig. 9 reveals
moments are aligned causing hysteresis with ideally zero that saturation of magnetization fails to occur even for the
remanence and coercivity [53, 54]. From the experimental applied magnetic field as high as 50 kOe. This behaviour
results, some multidomain structures with the single domain along with negligible coercivity and retentivity values can
size range exist due to disordered surface spins. The main also be attributed to SPM [57]. The maximum magnetisation
precursor for magnetic properties is exchange interactions at 5 K for A2 at 50 kOe is found to be 43 emu/g which is
between spins corresponding to different oxidation states. In much lower than the saturation magnetization (93.9 emu/g)
addition, the nanoparticles have a large volume to diameter for bulk cobalt ferrite. The lower values are due to surface
ratio value. Therefore, systems with noticeable hysteresis effects that become prominent at nanoscale such as spin
but lower Ms values also exist within the d­ SP range [55]. At canting and magnetically dead particle surface layer [6, 58]
low temperature, ferrimagnetic behaviour of cobalt ferrite and strong magnetic anisotropy [59, 60]. Structural defects
nanoparticles is dominant because of ordering of relaxing create lattice strain and influence the spin alignment. The
magnetic moments below the blocking temperature resulting surface area impacts upon the electron cloud density of the
in alignment of spins in the direction of applied magnetic cationic nuclei (Fe for Mössbauer). The nanoparticle size,
field. At room temperature, the thermal energy is sufficient cationic distribution and surface area are major factors that

13
Surfactant based synthesis and magnetic studies of cobalt ferrite Page 9 of 13  660

Fig. 7  a and b Particle size distribution from TEM results for A3 and B3, respectively

Fig. 8  a–d Mossbauer spectra for samples A2, A3, B2 and B3, respectively

13
660   Page 10 of 13 M. Gupta et al.

Table 2  Mössbauer parameters Sample Parameters I.S. (mm ­s−1) Q.S. (mm s­ −1) L.W. (mm ­s−1) Hhf (T) % Area (G)
Isomeric Shift (I.S.),
Quadrupole Splitting (Q.S.), A2 3+
Sx1Fe Tetra 0.33 0.07 1.33 441 36.7
and Hyperfine Field (Hhf)
Sx1Fe3+ Octa 0.34 0.08 0.74 481 45.8
calculated for samples A2–A3
and B2–B3 Db1 ­Fe3+Octa 0.34 0.61 0.5 – 17.5
A3 Sx1Fe3+Tetra 0.31 0.02 0.80 484 48.1
Sx1Fe3+ Octa 0.34 0.12 1.02 431 16.6
Db1 ­Fe3+Octa 0.34 0.75 0.53 – 35.3
B2 Sx1Fe3+Tetra 0.33 0.08 0.66 466 44.3
Sx1Fe3+ Octa 0.37 0.03 1.25 434 52.6
Db1 ­Fe3+Octa 0.41 0.57 0.3 – 3.1
B3 Sx1Fe3+Tetra 0.26 0.06 1.28 450 34.2
Sx1Fe3+ Octa 0.33 0.02 0.62 486 44.9
Db1 ­Fe3+Octa 0.33 0.69 0.65 – 20.9

Fig. 9  a–d M-H graph for samples A2-A3 and B2-B3 at 300 K and 5 K, respectively

influence the results from Mössbauer spectroscopy and mag- for very high values of the applied field, H. Here, Ms is the
netic measurements. saturation magnetisation and b is the fitting parameter. The
The saturation magnetisation ­(Ms) was measured by plot- intercept of the above straight line on Y-axis is the saturation
ting the function [61] magnetisation value. It is observed that the saturation mag-
netisation values increase with increasing surfactant concen-
M = Ms (1 − b∕H) (2) tration. This can be attributed to enhancement in crystalline

13
Surfactant based synthesis and magnetic studies of cobalt ferrite Page 11 of 13  660

structure with each sample and correlated with XRD and


Mössbauer results where the isomeric shift and quadrupole
splitting values are found to be closer to the expected range
for sample A3 [62]. From the graphs, SPM is more pro-
nounced in SDS mediated A series samples compared to
CTAB mediated B series. The particle sizes in both cases
are equivalent. However, from Mössbauer and magnetic
results, it can be concluded that the presence of spin disor-
der on the surface is greater for cobalt ferrite nanoparticles
synthesized using CTAB as surfactant because of higher
remanence and coercivity values of B series compared to
A series. The difference in ­Ms values for SDS and CTAB
systems is a combined effect of all these factors [63, 64]. The
ratio of remanence to saturation (Mr/Ms) or the squareness
ratio decreased with the increasing quantity of surfactant in
the system. The squareness ratio calculated for the samples
show a large deviation from the bulk value (0.50) which at
5 K is quite higher than the reported value and lower than
reported for room temperature measurement. Higher values
are obtained for A series at comparing Mr/Ms for samples A
and B. This is also the consequence of the superparamag-
netic nature of these nanoparticles to achieve saturation in
the given applied field.
The moment versus temperature measurements in the
presence of 100 Oe of the external magnetic field are in the
temperature range of 5 K to 300 K as shown in Fig. 10. The
upper portion of the M vs T curve refers to FC measurement
and the lower portion to ZFC, respectively. The point of
bifurcation of the FC and ZFC curves are blocking tempera-
tures that are found to be 300 K for all the samples.
Fig. 10  FC-ZFC plots for samples A2, A3, B2 and B3 at an applied The Law of Approach to Saturation (LAS) is used for
field of 100 Oe
calculating the anisotropy constant ­K1. For very high values
of the applied magnetic field (H) i.e. H >  >  > Hc (coerciv-
ity), magnetization M has been plotted as a function of H
(Fig. 11). The expression for LAS is [65]:
( ))
K1 2
(
8
M(H) = MS 1 − + 𝜒n H (3)
105 Ms H

Here, K1 is the cubic anisotropy constant and ratio 8/105


holds for random polycrystalline specimens with cubic
anisotropy. The term 𝜒n H is for spontaneous magnetiza-
tion occurrence known as forced magnetization observed
at high fields and 𝜒n is the high field susceptibility. In cubic
anisotropy, three prominent axes for anisotropy are usually
the principal axis. The axis of the applied magnetic field
determines the anisotropy axis for the given system. The
anisotropy constant has been calculated using the values of
applied magnetic field greater than 1 T and corresponding
magnetization values to plot M v/s H graph that was fitted
to Eq. (3) with K1 and Ms as dependents. The values are
Fig. 11  Approach to Saturation for sample A3 summarised in Table 3.

13
660   Page 12 of 13 M. Gupta et al.

Table 3  Magnetic measurement Sample Tempera- Hc (× 104 Oe) Mr (emu ­g−1) Ms (emu ­g−1) K1 × 105 ­(Jm−3) Mr/Ms
values using SQUID ture (K)

A2 5 1.5 28.4 43.0 11.6 0.66


300 0.01 2 32.7 3.52 0.06
A3 5 1.08 37.3 55.3 13.9 0.67
300 0.02 6 38.8 2.65 0.16
B2 5 1.0 38.4 60.79 6.88 0.63
300 0.11 20.6 49.34 2.38 0.42
B3 5 1.2 33.1 64.85 6.11 0.51
300 0.06 10.8 35.3 1.63 0.31

As can be seen from Table 3, anisotropy constant (K1) value is different for different surfactants. The purpose of
values for samples of both A and B series are lower than using ionic surfactants is to create spin distortions in the
for bulk values because of reduced size and thus reduced system and study the magnetic properties in such case. The
magnetostatic energy. The anisotropy constant has a larger results from Mössbauer are indicative of SPM. On com-
value at 5 K because of the absence of thermal fluctuations. paring SDS and CTAB synthesized samples, SPM is more
According to the Stoner-Wohlfarth theory [53], ­EA or dominant in SDS series. The CTAB synthesized particles
activation energy is a type of threshold for opposing change although within comparable sizes produce higher values of
in direction of magnetisation given by retentivity and coercivity than the SPM limit as a conse-
quence of unresolved spin on their surface. This causes a
EA = K1 V sin2 θ (4) few spins in non-uniform alignments giving higher M ­ r and
­Hc values as well as higher M
­ s. SDS synthesized cobalt fer-
where K1 is the anisotropy constant, V is the nanoparticle
rite nanoparticles have comparably ordered surface spins
volume and θ is the angle between the easy axis and applied
and narrow hysteresis making them susceptible to magnetic
field.
tuning for suitable applications.
When the particle size is decreased to beyond the criti-
cal diameter, EA is overcome by room temperature thermal Acknowledgements  The authors acknowledge the help received from
energy. As a result, SPM state of nanoparticles is reached UGC-DAE CSR, Kolkata Centre for providing the SQUID and Moss-
and the anisotropy also decreases. The coercivity values are bauer spectroscopy facilities and Centre for Research in Nanoscience
higher for the samples B2 and B3 compared to A2 and A3 and Technology, University of Calcutta for providing TEM facility. The
author (Meenal Gupta) is thankful to UGC for providing fellowship.
because of surface spin fluctuations. The magnetic response Prof. Datta is thankful to GGSIP University, New Delhi for the FRGS
for A series cobalt ferrite nanoparticles is superior to that grant GGSIPU/DRC/FRGS/2018/1/ (1115) and DST for FIST grant
of B series. It can be said that the magnetic morphology (SR/FST/PSI-167/2011(C)).
of the A series cobalt ferrite nanoparticles is better than
that of B series in terms of uniformity and stability in mag- Funding  GGSIP University, New Delhi for the FRGS grant GGSIPU/
DRC/FRGS/2018/1/ (1115) and DST for FIST grant (SR/FST/
netic results. As seen in Scheme 1, the impact of cationic PSI-167/2011(C)).
and anionic surfactants is different for same nanoparticle
synthesis because micelles formed by both surfactants are Compliance with ethical standard 
influenced by surface charges and counter ions that produce
either attractive or repulsive forces. These forces also influ- Conflict of interest  The authors declare that there is no conflict of in-
ence the micelle self-arrangement and hence the resulting terest regarding the publication of this paper.
nanoparticle assembly [40].

References
4 Conclusions
1. C. Kittel, Wiley Sons, New York, NY (2004).
2. D.S. Mathew, R.S. Juang, Chem. Eng. J. 129, 51 (2007)
The above work is a preliminary search for the best sur- 3. T.E. Quickel, V.H. Le, T. Brezesinski, S.H. Tolbert, Nano Lett.
factant to prepare highly crystalline cobalt ferrite and to 10, 2982 (2010)
determine the appropriate amount of the surfactant required. 4. C. Liu, A.J. Rondinone, Z.J. Zhang, Pure Appl. Chem. 72, 37
The cobalt ferrite material displays good crystalline struc- (2000)
5. W.H. Bragg, London, Edinburgh. Dublin Philos. Mag. J. Sci. 30,
ture as seen from XRD and TEM results for a particular 305 (1915)
amount of surfactant present in the microemulsion. This 6. R. Kodama, J. Magn. Magn. Mater. 200, 359 (1999)

13
Surfactant based synthesis and magnetic studies of cobalt ferrite Page 13 of 13  660

7. M. Sugimoto, J. Am. Ceram. Soc. 82, 269 (1999) 41. B. Payet, D. Vincent, L. Delaunay, G. Noyel, J. Magn. Magn.
8. R. Valenzuela, Phys. Res. Int. 2012, 1 (2012) Mater. 186, 168 (1998)
9. C. Cannas, A. Ardu, D. Peddis, C. Sangregorio, G. Piccaluga, A. 42. C.R. Vestal, Z.J. Zhang, Int. J. Nanotechnol. 1, 240 (2004)
Musinu, J. Colloid Interface Sci. 343, 415 (2010) 43. A. J. Rondinone, A. C. S. Samia, and Z. John Zhang, J. Phys.
10. M. Grigorova, H.J. Blythe, V. Blaskov, V. Rusanov, V. Petkov, V. Chem. B 104, 7919 (2000).
Masheva, D. Nihtianova, L.M. Martinez, J.S. Muñoz, M. Mikhov, 44. D. Makovec, A. Ko ak, and M. Drofenik, Nanotechnology 15,
J. Magn. Magn. Mater. 183, 163 (1998) S160 (2004).
11. P.C.R. Varma, R.S. Manna, D. Banerjee, M.R. Varma, K.G. 45. L. Yan, Y. Li, Z.-X. Deng, J. Zhuang, X. Sun, Int. J. Inorg. Mater.
Suresh, A.K. Nigam, J. Alloys Compd. 453, 298 (2008) 3, 633 (2001)
12. N.S. Gajbhiye, S. Bhattacharyya, G. Balaji, R.S. Ningthoujam, 46. C. N. Chinnasamy, B. Jeyadevan, K. Shinoda, K. Tohji, D. J.
R.K. Das, S. Basak, J. Weissmüller, Hyperfine Interact. 165, 153 Djayaprawira, M. Takahashi, R. Justin Joseyphus, and A. Naraya-
(2005) nasamy, Appl. Phys. Lett. 83, 2862 (2003).
13. N. Moumen, P. Veillet, M.P. Pileni, J. Magn. Magn. Mater. 149, 47. R.L. Mössbauer, Zeitschrift Für Phys. 151, 124 (1958)
67 (1995) 48. R. Bauminger, S.G. Cohen, A. Marinov, S. Ofer, E. Segal, Phys.
14. L.J. Cote, A.S. Teja, A.P. Wilkinson, Z.J. Zhang, Fluid Phase Rev. 122, 1447 (1961)
Equilib. 210, 307 (2003) 49. H.R. Rechenberg, F.A. Tourinho, Hyperfine Interact. 67, 627
15. T. Kodama, Y. Kitayama, M. Tsuji, Y. Tamaura, J. Magn. Soc. (1991)
Japan 20, 305 (1996) 50. E.C. Sousa, M.H. Sousa, G.F. Goya, H.R. Rechenberg, M.C.F.L.
16. P. Munnik, P.E. De Jongh, K.P. De Jong, Chem. Rev. 115, 6687 Lara, F.A. Tourinho, J. Depeyrot, J. Magn. Magn. Mater. 272–276,
(2015) E1215 (2004)
17. S. Zeng, K. Tang, T. Li, Z. Liang, D. Wang, Y. Wang, Y. Qi, W. 51. P.C. Morais, V.K. Garg, A.C. Oliveira, L.P. Silva, R.B. Azevedo,
Zhou, J. Phys. Chem. C 112, 4836 (2008) A.M.L. Silva, E.C.D. Lima, J. Magn. Magn. Mater. 225, 37 (2001)
18. H. Kavas, N. Kasapoğlu, A. Baykal, Y. Köseoğlu, Chem. Pap. 63, 52. O. Karaagac, B. Bilir, H. Kockar, J. Supercond. Nov. Magn. 28,
450 (2009) 1021 (2015)
19. J. Lee, J. Park, C. Kim, J. Mater. Sci. 3, 3965 (1998) 53. E.C. Stoner, E.P. Wohlfarth, IEEE Trans. Magn. 27, 3475 (1991)
20. C. Aubery, C. Solans, S. Prevost, M. Gradzielski, M. Sanchez- 54. A.E. Berkowitz, W.J. Schuele, J. Appl. Phys. 30, S134 (1959)
Dominguez, Langmuir 29, 1779 (2013) 55. C.P. Bean, J. Appl. Phys. 26, 1381 (1955)
21. N. Moumen, P. Bonville, M.P. Pileni, J. Phys. Chem. 100, 14410 56. Y. Il Kim, D. Kim, and C. S. Lee, Phys. B Condens. Matter 337,
(1996) 42 (2003).
22. C.T. Seip, E.E. Carpenter, C.J. O’Connor, V.T. John, and Sichu 57. A. Kale, S. Gubbala, R.D.K. Misra, J. Magn. Magn. Mater. 277,
Li. IEEE Trans. Magn. 34, 1111 (1998) 350 (2004)
23. K. Maaz, A. Mumtaz, S.K. Hasanain, A. Ceylan, J. Magn. Magn. 58. A.H. Lu, E.L. Salabas, F. Schüth, Angew. Chemie - Int. Ed. 46,
Mater. 308, 289 (2007) 1222 (2007)
24. E.W. Kaler, J.F. Billman, J.L. Fulton, R.D. Smith, J. Phys. Chem. 59. V. Blaskov, V. Petkov, V. Rusanov, L.M.M. Martinez, B. Martinez,
95, 458 (1991) J.S.S. Muñoz, M. Mikhov, M. Grigorova, H.J. Blythe, V. Blaskov,
25. D. Langevin, Annu. Rev. Phys. Chem. 43, 341 (1992) V. Rusanov, V. Petkov, V. Masheva, D. Nihtianova, L.M.M. Mar-
26. S.K. Mehta, S. Kumar, M. Gradzielski, J. Colloid Interface Sci. tinez, J.S.S. Muñoz, M. Mikhov, V. Rusanov, L.M.M. Martinez,
360, 497 (2011) B. Martinez, J.S.S. Muñoz, M. Mikhov, J. Magn. Magn. Mater.
27. S. Kumar, V.K. Aswal, J. Kohlbrecher, Langmuir 28, 9288 (2012) 162, 331 (1996)
28. S. Paria, K.C. Khilar, Adv. Colloid Interface Sci. 110, 75 (2004) 60. Y. Ahn, E.J. Choi, S. Kim, H.N. Ok, Mater. Lett. 50, 47 (2001)
29. B.D. Cullity, J.W. Weymouth, Am. J. Phys. 25, 394 (1957) 61. H.H. Hamdeh, J.C. Ho, S.A. Oliver, R.J. Willey, G. Oliveri, G.
30. M. Wu, J. Long, A. Huang, Y. Luo, S. Feng, R. Xu, Langmuir 15, Busca, J. Appl. Phys. 81, 1851 (1997)
8822 (1999) 62. S.M. El-Sheikh, F.A. Harraz, M.M. Hessien, Mater. Chem. Phys.
31. A.E. Palmqvist, Curr. Opin. Colloid Interface Sci. 8, 145 (2003) 123, 254 (2010)
32. M. A. López-Quintela, C. Tojo, M. C. Blanco, L. García Rio, and 63. A.H. Morrish, S.P. Yu, Phys. Rev. 102, 670 (1956)
J. R. Leis, Curr. Opin. Colloid Interface Sci. 9, 264 (2004). 64. R.H. Kodama, A.E. Berkowitz, E.J. McNiff Jr., S. Foner, Phys.
33. H. Khatri, G. Packiaraj, R.B. Jotania, Adv. Mater. Res. 938, 14 Rev. Lett. 77, 394 (1996)
(2014) 65. H. Kronmüller, M. Fähnle, M. Domann, H. Grimm, R. Grimm, B.
34. H. Kunieda, A. Nakano, M.A. Pes, Langmuir 11, 3302 (1995) Gröger, J. Magn. Magn. Mater. 13, 53 (1979)
35. K. Aramaki, T. Hayashi, T. Katsuragi, M. Ishitobi, H. Kunieda, J.
Colloid Interface Sci. 236, 14 (2001) Publisher’s Note Springer Nature remains neutral with regard to
36. G. Schmid, Prog. Colloid Polym. Sci. 111, 52 (1998) jurisdictional claims in published maps and institutional affiliations.
37. M.A. López-Quintela, Curr. Opin. Colloid Interface Sci. 8, 137
(2003)
38. D. V Leff, P. C. Ohara, J. R. Heath, and W. M. Gelbart, 7036
(1995).
39. E. Matijević, Langmuir 2, 12 (1986)
40. Z.Y. Shen, M.T. Lee, Materials (Basel). 10, 555 (2017)

13

You might also like