Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

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

net/publication/321033044

Ferrites Obtained by Sol–Gel Method

Chapter · October 2017


DOI: 10.1007/978-3-319-19454-7_125-3

CITATIONS READS

48 5,931

5 authors, including:

Sagar E. Shirsath Danyang Wang


UNSW Sydney UNSW Sydney
303 PUBLICATIONS   8,490 CITATIONS    91 PUBLICATIONS   1,470 CITATIONS   

SEE PROFILE SEE PROFILE

Swati Jadhav Maheshkumar L. Mane


Yadavrao Taskgaonkar Institute of Engineering & Technology Shikshan Maharshi Guruvarya R G Shinde Mahavidyalaya Paranda
18 PUBLICATIONS   512 CITATIONS    67 PUBLICATIONS   1,800 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Room Temperature Crystallization and Magnetic Anisotropy View project

Synthesis Characterization and studies on magnetic properties of Magneto electric Nanocomposites View project

All content following this page was uploaded by Sagar E. Shirsath on 28 September 2019.

The user has requested enhancement of the downloaded file.


Ferrites Obtained by Sol-Gel Method
25
Sagar E. Shirsath, Danyang Wang, Santosh S. Jadhav,
M. L. Mane, and Sean Li

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 696
Sol-Gel Method: Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
Powder Ferrite Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 699
Nanotubes and Nanowires . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705
Nanoribbon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 707
Ferrite Thin Film . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 709
Spinel Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 711
Hexaferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 718
Bismuth Ferrite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 732

Abstract
Ferrites are ferrimagnetic materials, a broad class of magnetic oxides with its
remarkable structural, electrical, and magnetic properties making them suitable
for a variety of applications. These properties strongly depend upon the method of
preparation. Sol-gel method is having countless advantages over other methods to

S. E. Shirsath (*) · D. Wang · S. Li


School of Materials Science and Engineering, University of New South Wales, Sydney, NSW,
Australia
e-mail: shirsathsagar@hotmail.com; s.shirsath@unsw.edu.au; dy.wang@unsw.edu.au; sean.
li@unsw.edu.au
S. S. Jadhav
Department of Physics, Dnyanopasak Shikshan Mandal’s Arts, Commerce and Science College,
Jintur, India
e-mail: santosh.jadhav28@yahoo.com
M. L. Mane
Department of Physics, SGRG Shinde Mahavidyalaya, Osmanabad, Maharashtra, India
e-mail: mane.maheshkumar@hotmail.com

# Springer International Publishing AG, part of Springer Nature 2018 695


L. Klein et al. (eds.), Handbook of Sol-Gel Science and Technology,
https://doi.org/10.1007/978-3-319-32101-1_125
696 S. E. Shirsath et al.

obtain the ferrite with desired nanoarchitecture. In this chapter, the effects of
varying synthesis conditions such as chelating agent/fuel to nitrate ratio, effect of
pH, sintering time/temperature, etc., on the final product of ferrites are discussed.
This chapter also focuses to synthesize the ferrite in the form of nanocrystalline
powder, nanowires, nanotubes, nanoribbon, mesoporous material and thin film of
spinel, hexaferrite, and perovskite-bismuth ferrite. Possible mechanisms that
affect the structure formation and properties of ferrite along with their application
in various technological fields are also discussed.

Introduction

Several decades ago, research efforts were focused on active solids, which are of
great scientific and technological importance. Ferrites consist iron oxide as their
main component possess class of magnetic oxides materials showing remarkable
structural, electrical, dielectrical, and magnetic properties that have been studied
over the last 70 years (Neel 1948). The exceptional properties of ferrites include
significantly high saturation magnetization, a high electrical dc resistivity, consider-
ably low electrical losses, with a high chemical stability. They are extensively used
in various applications starting from simple lifting magnets to the most complex
microwave communication system employed in the outer space. They can be used in
satellite communication, radar, memory of computer, radio, television, videotape,
etc. The most efficient power supplies used in electronic appliances are the switch
mode power supplies (SMPS), which have ferrite as a major part.
Magnetic resonance imaging (MRI) enhancement, catalysis, sensors, and pig-
ments (Mathew and Juaug 2007) are the few of the developing areas of the
applications of ferrites. These novel applications of the ferrites are possible due to
the synthesis of the ferrites at nanoscale. Thus, nanoscale size of the spinel ferrites
has opened up the door for intensive research to utilize their properties for different
novel applications. It includes applications of the ferrites as the material for
magnetocaloric effect (MCE), sensors, biosensors, supercapacitors, etc. (Gopalan
et al. 2010; Gass et al. 2008; Darshane and Mulla 2010; Haun et al. 2010).
Ferrites, the magnetic ceramics consisting of Fe3+ as the major constituent, were
initially prepared by a solid-state reaction mechanism known as standard ceramic
method. This synthesis method involves stoichiometry mixing and grinding (for
long hours) powders of the constituent oxides chemicals, carbonates, and heating
them at relatively high temperature (up to ~1000  C) with intermediate grinding, if
necessary. The ceramic method is thus called as a time-consuming and a brute force
method. It also yields particles of comparatively large course because of high-
temperature treatment. One of the other major drawbacks is the irregularity in
shape, size, and stoichiometry, which may result in the nonreproducible ferrite
products. To overcome these drawbacks of standard ceramic method, numerous
chemical methods were explored (Rao 1993) including wet chemical coprecipitation
method (Jadhav et al. 2010; Patange et al. 2011), sol-gel methods (Gore et al. 2015;
Shirsath et al. 2014), thermal decomposition method (Vestal and Zhang 2003),
25 Ferrites Obtained by Sol-Gel Method 697

micelle (Sharifi et al. 2012) and reverse micelle method (Thakur et al. 2009), Pechini
method (Pechini 1967), etc. Chemical synthesis approaches have the advantage over
the ceramic method because (i) it requires low temperatures, (ii) it yields fine
particles of nanosize, (iii) the method is easy and cheap and requires less time,
(iv) stoichiometry does not change, (v) and importantly the product is reproducible.
Chemical method essentially makes use of simple reactions such as inter-
collation, ion-exchange, hydrolysis, dehydration, and reduction that can be carried
out at relatively low processing temperature. Developing a method of synthesis of
ferrites with optimum properties is difficult and complex as well. The main problem
is induced since most of the properties requirement for ferrite applications are
extrinsic and not intrinsic, where the ferrite is defined by its chemistry and crystal
structure, as well as requires knowledge and control of different parameter of its
microstructures, such as bulk density, grain/crystallite size, percentage porosity, and
their intra- and inter-granular distribution. The chemical synthesis methods cited
above are few among the methods suitable in these criteria. In the area of magnetism
of fine particles, modifications in saturation magnetization and magnetic anisotropy
have been affected by the reduction of particle and grain sizes. The corresponding
changes in magnetism of surface layers have been the focus of current interest.
Research in the field of fine-grain ceramic oxide systems has gained immense
importance because of their potential applications in many areas of technology.
Among the available chemical methods, sol-gel synthesis is known to be a
versatile technique that gives rise to a new class of novel materials by enabling the
synthesis of nanoparticles. The chemical combination and crystallography of these
materials are analogous to their alloys prepared by the traditional methods like solid-
state reaction routes, and possess different but better morphology owing to their fine
particle size. The new investigations in sol-gel technology make it possible to
develop the method of crystallization in the colloidal state at relatively low temper-
atures ~100  C (Toksha et al. 2011). This opened a door to the synthesis of
nanocrystalline oxides, mixed oxides, and the composite materials prepared by
low-temperature sol-gel method.

Sol-Gel Method: Chemistry

Figure 1 displays different steps involved in material synthesis by sol-gel method. As


the name indicates, the sample in this method to be synthesized must transit through
two phases, viz., “sol” and “gel.” “Sol” is a colloidal suspension of solid particles in
a liquid media. In colloidal suspension, the interactions are governed by van der
Waals forces and the particles perform Brownian motion in the liquid. Such solution
is used to generate polymers or particles from which ceramic materials can be
prepared. The colloidal sol is prepared using the chemicals where the metals or
metalloids are surrounded by ligands, e.g., nitrates of corresponding metals. The
salts like nitrates are easily soluble in water. The reaction is called hydrolysis in
which hydroxyl ion attached to the metal ion (M), as in the reaction,
698 S. E. Shirsath et al.

a b
Solution Dip coating Spray drying

Sol
Powders Films Electro
Sol deposition
+ - Electrospinning

Patterns
Fibres

Flow coating
Gel
Patterns
Blow spinning

Spin Coating
Monoliths
Powders

Pulling Pulling
down up

Fig. 1 Material synthesis by sol-gel method; (a) different forms of synthesized materials; (b)
sol-gel methods to synthesize different forms of materials (Sanchez et al. 2011)

MðORÞ4 þ H2 O ! HO  MðORÞ3 þ ROH (1)

Depending on the amount of water and catalyst present, the hydrolysis completes
with all OR groups replaced by OH,

MðORÞ4 þ 4H2 O ! MðOHÞ4 þ 4ROH (2)

or the hydrolysis stops while the metal (M) is only partially hydrolyzed, M
(OR)4x(OH)x. The two such molecules may link together through condensation
reaction, such as

ðORÞ3 M  OH þ HO  MðORÞ3 ! ðORÞ3 M  O  MðORÞ3 þ H2 O (3)

The condensation reaction thus liberates a water molecule to build a larger metal-
containing molecule through the process of polymerization. A polymer is a huge
molecule formed by hundreds or thousands of small molecules called monomers that
are capable of forming at least two bonds. If these monomers make more than two
bonds, then it results in the formation of unlimited size of a giant macroscopic
molecule. As the molecule reaches to macroscopic dimensions extending throughout
the solution, the substance formed is termed as a “gel.” Thus, gel is a stage of
substance which contains a continuous solid structure (a giant macroscopic
25 Ferrites Obtained by Sol-Gel Method 699

molecule) enclosed in a continuous liquid phase that possesses elasticity. The gels
are mostly amorphous, i.e., noncrystalline in nature but can be crystallized upon heat
treatment. Whenever our aim is to prepare a final product like a ferrite or any other
ceramic materials, it is necessary to heat the gel at a temperature high enough to
optimize the sintering.

Powder Ferrite Nanoparticles

In the sol-gel method usually recognized as sol-gel autocombustion or sol-gel


autoignition or sol-gel self-combustion, the nitrates of the constituent metal ions
and a suitable chelating agent (fuel), such as urea/glycine/citric acid, are used as the
starting materials, owing to their easy solubility in water.
The flowchart of the sol-gel method is shown in Fig. 2. The metal nitrate to citrate
ratio is generally kept as 1:x (where, x could be 1 to 3 in most of the cases). Reaction
procedure is generally being carried out in air atmosphere and without the protection
of any inert gases. The metal nitrates should be dissolved together in a minimum
amount (say 100 ml for 10 gm of metal nitrates) of double-distilled or deionized
water to obtain a clear solution.
An aqueous solution of citric acid needs to mix with metal nitrates solution that
should be placed onto a hot plate (at 100  C) with continuous stirring. The solution
generally became viscous that forms a brown gel. Viscous gel began frothing after
the evaporation of all water molecules from the mixture. In the later stage, the gel
automatically ignited and burnt with high temperature glowing flints. The ignition

Fig. 2 Flowchart of the Nitrates Citric Acid


sol-gel autocombustion
method

MN:CA = 1:x
Aqueous solution

Ammonia
Mixed solution
Stirring at ∼90°C
Sol

Gel

Combustion

Ferrite
700 S. E. Shirsath et al.

O OH
O- O- O-
-
-
O N+ O N+ -
O N+ HO
-
Co++
O- O N+ Cr+++
O- -
O N+ Fe+++ O- HO OH
O N+ O- N+ O- N+ O
O- O- O- O
Cobalt nitrate chromium nitrate (x) ferric nitrate (2 - x) citric acid

Heating and NH3


Stirring ammonia

Heating and O-- O--


Stirring O Fe+++
--
O--
Co++
O-- O Cr+++
--
O--
CoCr(x)Fe(2-x)O4 +++ Fe +++
cobalt oxide Cr
chromium oxide (x) ferric oxide (2 - x)

Fig. 3 The chemical reaction of the ferrite series CoCrxFe2xO4 synthesized by sol-gel auto-
combustion method

and decomposition process continues till the consumption of whole citrate complex.
The autoignition can be completed within few minutes, yielding the brown-colored
ashes termed as a precursor. Such a chemical reaction for chromium-substituted
cobalt ferrite is shown in Fig. 3.
The ratio of fuel to nitrates can highly affect the particle size and magnetic
properties. The oxidizer to fuel mixture ratio can be computed in terms of the
elemental stoichiometric coefficient (Patil et al. 2008). Effect of the chelating/
combustion/fuel agents on the properties of Zn ferrite was reported (Slatineanu
et al. 2012). Six chelating/combustion agents, viz., citric acid, glucose, tartaric
acid, egg white, glycine, and urea, were used to obtain ZnFe2O4 nanoparticles by
a sol-gel autocombustion method. They observed that nature of fuel strongly affects
the autocombustion reaction intensity and the duration of thermal treatment, which
dictates specific structural characteristics and magnetic properties of the samples to
be used in biomedical applications.
Synthesis of NiFe2O4 (NFO) by sol-gel method uses different fuel/chelating
agents, viz., citric acid (C8H12O7) (CA), cellulose (C6H10O5) n (CL), tartaric acid
(C4H6O6) (TA), glycine (C2H5NO2) (G), cellulose–citric acid mixture (CL–CA),
hexamethylenetetramine (C6H12N4) (HA), and urea (CH4N2O) (U). Ni(NO3)26H2O
and Fe(NO3)39H2O were used as cation sources. The atomic ratio of metal cations
Ni2+ and Fe3+ is kept as 1:2, and the ferrite–fuel/chelating agent molar ratio is 1:3.
Cellulose was used with a ferrite/CL mass ratio = 1:1.8. A mixture of CL–CA with a
ferrite/CL/CA mass ratio = 1:1.8:2.7 was used as fuel/chelating agent. Table 1
shows flame temperatures measured with a UT infrared thermometer the during
auto-combustion for all fuel/chelating agents (Dumitrescu et al. 2013).
Fuel/chelating agent govern the combustion process. In the case of hexamethy-
lenetetramine and glycine, the process is spontaneous and violent, releasing large
25 Ferrites Obtained by Sol-Gel Method 701

Table 1 Flame Fuel agent Flame temperature ( C)


temperature for all
Citric acid (CA) 818
chelating/fuel agents used
in synthesis (Dumitrescu Tartaric acid (TA) 505
et al. 2013) Cellulose (CL) 650
Cellulose/citric acid (CL–CA) 700
Glycine (G) 1200
Hexamethylenetetramine (HA) 1100
Urea (U) 600

quantities of gas (nitrogen and carbon dioxide) and heat where the flame temperature
rises above 1000  C (Table 1). When citric acid, cellulose, tartaric acid, and
cellulose–citric acid mixture are used as fuel/chelating agents, the process takes
longer time (approximately over 30 min), whereas it is done within 2 min during the
use of urea. The arrangement of metal cations within the crystal structure is respon-
sible for such results that influenced the occurrence of combustion stage. Since both
hexamethylenetetramine and glycine induce an extremely violent combustion, cat-
ion arrangement within the spinel lattice occurs more quickly. On the contrary, citric
acid, tartaric acid, and urea induce slow combustions; therefore, cation arrangement
within the crystal lattice occurs comparative slowly. Combustion process slows
down considerably in the case of cellulose and cellulose–citric acid mixture,
resulting into cationic rearrangement that occurs on the existing cellulose micro-
structure. Therefore, while using cellulose and cellulose–citric acid mixture, crys-
tallites grow on the cellulose microfibers. Atom diffusion followed by metal cations
redistribution occurred in close proximity during the combustion and ferrite particle
formation process. In the quick combustion process, the energy released and the
process duration are not sufficient to allow diffusion and migration of atoms in the
crystal lattice at longer distances. The quick release of high amount of gas during the
combustion process denied the interactions among the constituent metal ions of the
crystal lattice. Pure spinel monophase can only be achieved for the NiFe2O4–glycine
sample, after the post-annealing at 500  C. This may be due to the extremely violent
combustion reaction that occurred during the synthesis with glycine as a fuel – the
mixture reaching a very high temperature (1200  C). A rapid release of large volume
of gas and temperature favors crystallite growth and the formation of single phase
spinel structure. Other fuel requires high temperature post-annealing to obtain the
single spinel phase. It is to be noted that crystallite and grain size are significantly
governed by the type of fuel/chelating agent and temperature used during synthesis.
This also influenced the catalytic activity of ferrite (Dumitrescu et al. 2013).
NiFe2O4–glycine sample with its pure spinel structure obtained by post-annealing
at 500  C do not show significant catalytic activity. Further, it has been observed that
the size of the particles depends on the use of fuel/chelating agents, leading to the
reduction of the contact surface that affect the catalytic activity of ferrites. The
existence of the α-Fe2O3 secondary phase enhances the catalytic activity by improv-
ing the intimate interactions among the two phases (Florea et al. 2009). The
702 S. E. Shirsath et al.

existence of α-Fe2O3 phase resulted in the increment of Fe3+ ions in the octahedral
sites, causing the reduction reaction of Fe3+ to Fe2+. Eventually, reduction reaction
leads to the release of a large volume of oxygen (Albuquerque et al. 2012). A
possible mechanism of hydrogen peroxide decomposition in the presence of
NiFe2O4 obtained is expressed in the following manner:

Fe2þ þ HO  OH ! Fe3þ þ HO þ HO •


Fe3þ þ HO  OH ! Fe2þ þ Hþ þ HOO •
HOO • þ HO • ! O2 þ H2 O
1
HO • þ • OH ! O2 þ H2 O
2
Hþ þ HO ! H2 O

In the sol-gel combustion synthesis, urea as a fuel affects the reaction and
properties of ferrite (Costa et al. 2001). The nitrates were mixed with urea, which
varied from the stoichiometric composition (without an excess of urea) up to
compositions of 40–100 wt percent of excess urea content. Increasing the urea
content gave rise to longer reaction flame times, altering the characteristics of the
prepared sample. The composition without excess urea showed the best results in
function of the characteristics of the prepared sample. Various fuels like ethylene
glycol (Gopalan et al. 2010; Li 2011), tetrakis (2-hydroxyethyl) orthosilicate
(Gharagozlou 2011), citric acid (Banerjee et al. 2011), acetic acid, carbamide and
acrylic acid (Hu et al. 2011), citric acid and ethylene glycol (Kumar and Yadav
2011), urea and ammonium nitrate (Deraz 2011), egg white (Hou et al. 2011),
modified sol-gel technique with aloe vera plant-extracted solution (Laokul et al.
2011), etc., have also been reported in the literature.
Solution-based sol-gel synthesis method is sensitive to amount of pH that deter-
mines the formation of the desired compound as well as its combustion process. A
study of CoDy0.1Fe1.9O4 ferrite synthesized by the sol-gel autocombustion method
shows that the combustion rate is significantly affected by the amount of pH in the
mixed solution (Shirsath et al. 2013). pH of the solution is controlled by the amount
of ammonia (NH3). With an increase in pH value from 2.5 to 7.5, the combustion rate
increased and thereafter decreased for pH = 10. The best possible results observed at
a pH = 7.5, where the water molecules evaporated from the solution and the viscous
gel was ignited in the air where the combustion rapidly propagated till the entire gel
burnt out to form final product in the form of a brown powder. In this case, the
decomposition/combustion reaction continued until the entire citrate complex was
consumed. Thermogravimetric (TG) study evidenced minor differences in the TG
curves below 250  C. The decreasing amount of H2O trapped inside the powder
resulted in decrease of weight loss with the increase in pH value from 2.5 to 7.5. In
fact, the increase of the concentration of NH3 ion leads to more metal complexes
polymerization and therefore to a less open structure that can accommodate at lower
percentage of H2O. Around this temperature range, some of the NH4NO3 formed
during synthesis may liberate NOx and O2 gas through following the decomposition
reaction:
25 Ferrites Obtained by Sol-Gel Method 703

Fig. 4 X-ray diffraction patterns and TEM images of CoDy0.1Fe1.9O4 nanoparticles at different pH
values (Shirsath et al. 2013)

NH4 NO3 ! NOx þ O2 þ N2 þ H2 O (4)

The oxygen gas can speed up the combustion process, and a large amount of heat
can be generated through exothermic reaction (Yue et al. 2004). The removal of
residuals (gel particulates, NH3, and trapped gases) can result into weight loss
through burnt powders during heating. The difference between the weight loss for
each powder is co-related with its combustion activity during synthesis. That support
the lowest weight loss for pH 7.5 sample because of the highest combustion rate
during its synthesis.
The low-intensity Co3O4 phase besides the α-Fe2O3 and γ-Fe2O3 phases appeared
in the XRD pattern pH 2.5 sample. In this case, nominal ratio of Fe/Co was 1.9:1; the
separation of a low quantity of cobalt oxide in pH 2.5 samples leads to a cobalt
deficient ferrite (Fig. 4). Adding NH3 into the aqueous solution of metallic nitrates
and citric acid promoted the metallic ions to be chelated by carboxylic groups
(COO) from citric acid in the solution, more completely. At pH 7.5, sufficient
carboxylic groups ionized from citric acid were available to completely chelate the
Fe3+, Dy3+, and Co2+ ions, resulting in the formation of the solid precursor with
uniformly distributed Fe3+, Dy3+, and Co2+ ions. Highest combustion rate during the
pH 7.5 resulted in the higher particle size and bulk density that also evidenced by the
lowest weight loss for this pH value.
The nanoparticles synthesized at pH 2.5 exhibit greater agglomerations as com-
pared to the products synthesized with other pH values (Fig. 4). It is a known fact
that the agglomeration among the particles is related to shape factor, porosity, surface
area, bulk density, etc. It is to be noted that most of the colloidal particles are
electrically charged, e.g., most metal oxides have a surface layer of the metal
hydroxide which is amphoteric and can become either positively or negatively
charged, by taking up a proton or by proton abstraction, depending on the amount
of pH. The electrostatic potential on the particle surface, relative to the surrounding
fluid, is largely dependent on the balance among the positive (H+) and negative
(OH) ions. These are the potential-determining ions in colloids and oxide system
704 S. E. Shirsath et al.

where the surface charge and potential are governed largely by the balance among
H+ and OH in solution, i.e., by the pH (Shirsath et al. 2013; Rashad et al. 2009).
Structural and morphological investigations of the comparative studies on the
catalytic activity of cobalt ferrite and bismuth-substituted cobalt ferrite nanoparticles
prepared by two different techniques, viz., combustion and coprecipitation, showed
that properties are independent of method of synthesis (Kiran and Sumathi 2017).
Among the predicted methods of preparation, ferrite nanoparticles synthesized by
combustion method show good catalytic activity than coprecipitation method.
CuFe2O4 (copper ferrite) nanoparticles were synthesized by the sol-gel method
and microwave method using sucrose as a fuel. The use of sucrose resulted in
shortened reaction time for microwave method and vice versa for sol-gel method
(Raja et al. 2016).
Aluminum (Al3+)-substituted nickel ferrite (NiFe2O4) nanoparticles were synthe-
sized by the chemical coprecipitation and sol-gel technique (Gul and Pervaiz 2012)
and investigated the structural, morphological, and electrical properties. The samples
synthesized by sol-gel method are comparatively more uniform and homogeneous,
narrowly distributed in size, with a smaller in particle size as compared to
coprecipitation method.
The structural, morphological, and magnetic properties of x%
(Ni0.65Zn0.35Fe2O4)/(100  x) % SiO2 ferrimagnetic nanocomposites for lower
concentration of Ni0.65Zn0.35Fe2O4 ferrite were synthesized by modified sol-gel
method (Stoia et al. 2011). The improvement consists in using an excess of ethylene
glycol (corresponding to a 1:1 M ratio TEOS:EG) and applying the post-annealing at
573 K for 3 h on the obtained gels. Low ferrite concentration and the annealing
temperature of the precursor gels influenced the structure, morphology, and the
magnetic behavior of the Ni0.65Zn0.35Fe2O4 ferrite nanoparticles that are embedded
in silica matrix.
Szczygiel and Katarzyna (Szczygiel and Winiarska 2011) have prepared
Mn0.6Zn0.4Fe2O4 by two-stage synthesis route. Here, a precursor for sol-gel auto-
combustion method was prepared by the co-precipitation technique. The sintering of
the as-synthesized product in an air atmosphere at a higher temperature favors the
formation of phase α-Fe2O3, which above 1000  C again dissolves in the spinel
phase. A pure spinel phase was only obtained by sintering the pressed product at
1300  C, resulted into well-densified homogeneous microstructure with average
smaller grain size compared product synthesized by conventional method.
Exchange bias, memory, and freezing effects in NiFe2O4 nanoparticles embedded
in SiO2 matrix were synthesized sol-gel method (Nadeem and Krenn 2011). The
particle size varies in the range of 8–12 nm and considered to be hybrid particles at
low temperatures with a core–shell structure. These core-shell structures mediate the
exchange bias among frozen surface spins and giant core spin.
The dielectric behavior of nanocrystalline NixZn1xFe2O4 (0  x  1) was
studied by Bhattacharjee et al. (2011). During the synthesis citric acid (C6H8O7)
and ethylenediamine (C2H8N2) were used as a coordinating agent and bridging
ligand, respectively. The new observation of nonlinear variation of the frequency-
dependent dielectric permittivity with respect to composition has been explained
25 Ferrites Obtained by Sol-Gel Method 705

successfully. S. Nasir et al. (2011) shows that nanoferrites can be synthesized by


simplifying sol-gel method without the use of water and surfactants for the study of
structural, dielectric, and electrical properties of zinc-doped nickel ferrite. In this
method, water and surfactants are avoided which result in higher purity of the
synthesized powder.
A two-step sintering process, first in air and subsequently in Ar (argon), helps to
retain particle size and shape of the NiFe2O4 nanoparticles networked (Waqas and
Qureshi 2010). In this method, sol-gel-derived nickel-iron oxide aerogels are trans-
formed into monodisperse, networked nanocrystalline magnetic oxides of nickel
ferrite with particle diameters that can be ripened with increments in temperature
under Ar atmosphere to 4.6, 6.4, and 8.8 nm. Here they show that the atmosphere
and processing temperature parameters govern critical roles in controlling the
crystallization of nickel ferrite nanoarchitectures.

Nanotubes and Nanowires

In recent years, considerable progress in synthesizing one-dimensional


(1D) magnetic nanostructures, such as nanofibers, nanotubes, and nanoribbons,
has been inspiring many researchers to investigate their magnetic-domain configu-
rations and magnetization reversals and to develop their electromagnetic nanoscale
device applications. Moreover, 1D magnetic nanostructures possess remarkable
geometrical limitations being comparable to the critical magnetic lengths such as
exchange length and domain wall width and can effectively overcome the serious
aggregation of nanoparticles. It means that 1D permanent magnetic nanostructures
may get more remarkable magnetocrystalline anisotropy and shape anisotropy (Sun
et al. 2005) that make them capable of offering more creativity for magnetic media,
especially the perpendicular magnetic recording media.
Coaxial nanostructures exhibit additional remarkable effects and have potential
applicability as multifunctional materials. Therefore, these materials have attracted
considerable attention from the point of view of both technological as well as
fundamental interests. Permalloy (Ni80Fe20) and multiferroic (BiFeO3,
BFO/BiFe0.95Co0.05O3, BFC) materials formed hybrid core–shell nanostructures
were synthesized by wet chemical impregnation followed by electrodeposition
within porous alumina membranes (Javed et al. 2015). Here the nanotubes of both
BiFeO3/BiFe0.95Co0.05O3 were designed by sol-gel method. Nanoporous anodic
aluminum oxide (AAO) templates with pore size of 300 and 100 nm were then
dipped into the sol for 20 min. A subsequent post-annealing of the templates
containing the precursors was carried out at 650  C, resulting in the formation of
nanotubes in the pores of the AAO templates. Figure 5 shows the schematic for
synthesis of BFO and NiFe nanotubes.
Figure 6 shows the SEM images of BiFe0.95Co0.05O3 and Ni80Fe20–(BiFeO3/
BiFe0.95Co0.05O3) core–shell nanowires of 300 and 100 nm, respectively. Further,
ferroelectric and magnetic study confirmed the coexistence of spontaneous electric
polarization and antiferromagnetic spin ordering in the BiFeO3 nanomaterials.
706 S. E. Shirsath et al.

Fig. 5 Ni80Fe20–(BiFeO3/BiFe0.95Co0.05O3) core–shell nanostructures prepared by template-


assisted sol-gel and electrodeposition by two-step method (Javed et al. 2015)

Fig. 6 SEM image of (a) BFC nanotubes with a diameter of 300 nm and (b) NiFe-BFC core–shell
nanowires with diameter 100 nm (Javed et al. 2015)

Ferromagnetic cores (Ni80Fe20) are exchange coupled with the multiferroic BiFeO3
shell. Exchange bias can be found in the Ni–BFO core–shell nanostructures without
employing the magnetic-field annealing. Ferromagnetic nanowires/nanotubes lead to
the formation of core–shell nanostructured ferromagnetic–multiferroic composite.
By doping Co2+ in BiFeO3 and using Ni80Fe20 as core (compared with Ni–BFO),
25 Ferrites Obtained by Sol-Gel Method 707

a significantly enhanced exchange bias field (Hex) and improved Hc have been
observed. The rearrangement of the antiferromagnetic domains during the electro-
deposition of Ni80Fe20 nanowires/nanotubes inside the BiFeO3/BiFe0.95Co0.05O3
shell resulted in enhanced Hex for Ni80Fe20–(BiFeO3/BiFe0.95Co0.05O3) core–shell
nanotubes. Therefore, ferromagnetic Ni80Fe20 cores are exchange coupled to the
multiferroic BiFeO3/BiFe0.95Co0.05O3 shell that can be enhanced by doping of
appropriate ferromagnetic/multiferroic material. With this ferromagnetic/multi-
ferroic 1D nanostructures with magnetic-field control of polarization or electric-
field control of magnetization can be studied (Javed et al. 2015).

Nanoribbon

Electrospinning technique is more simple and flexible for producing continuous 1D


nanostructures of various materials by using a high-voltage dc source. Moreover,
nanoribbons also could be considered as a development by cutting a finite-width
slice from the 2D nanosheets. In this context, width-controlled SrFe12O19 (strontium
ferrite) nanoribbons were synthesized via a polymer sol-assisted single-spinneret
electrospinning route followed by heat treatment in air (Jing et al. 2015).
Firstly, 0.018 g of anhydrous Sr(NO3)2 (strontium nitrate) and 0.368 g of Fe
(NO3)39H2O (ferric nitrate) were easily dissolved in 1.5 g of deionized water.
Subsequently, a certain amount of polyvinylpyrrolidone (PVP) powders (0.4, 0.5,
and 0.6 g as needed) and 2.4 g C2H5OH were added into the above red nitrate
solution under vigorous stirring and equilibrated for overnight to acquire a homo-
geneous viscous solution. Of which, the PVP concentrations of these obtained
solutions were about 8.5%, 10.4%, and 12.3%, respectively. Secondly, proper
amount of the obtained spinning solution was transformed into a glass syringe
equipped with a stainless needle of an inner diameter was about 0.4 mm for
electrospinning. The needle was then connected with a positive voltage of 15 kV
and the aluminum collector was grounded. The vertical distance between the tip of
the needle and the collector plane was kept at ~20 cm. Moreover, the spinning
solution was withdrawn at a rate of 0.3 mL/h by a microinjection pump. The whole
electrospinning process was conducted at room temperature (about 25  C) in air.
Thirdly, the collected precursor PVP/SrFe12O19 nanoribbons were kept in a drying
oven for several hours and then were subjected to annealing at 800  C for 2 h in a
muffle furnace in air atmosphere. The heating and cooling rates were both 1  C/min.
In this way, SrFe12O19 nanoribbons can be synthesized.
To obtain a visual understanding for the formation mechanism of SrFe12O19
nanoribbons, a possible schematic diagram is displayed in Fig. 7. The whole
preparation can be segmented into electrospinning and heat treatment. Based on
the earlier researches, it can be summarized that the novel ribbon-like structure of
SrFe12O19 nanoribbons is produced during the electrospinning process. When the
spinning solution arrived up to the spinneret tip, a Taylor cone-shaped colloidal
droplet (Fig. 7a) is formed under the coaction of electrostatic field force and surface
tension. With the accumulation of surface charges, the electrostatic field force
708 S. E. Shirsath et al.

Fig. 7 Schematic diagram of the formation mechanism of SrFe12O19 nanoribbons (Jing et al. 2015)

eventually overcomes the surface tension and a columnar flow jet (Fig. 7b) with a
circular cross section (Fig. 7c) ejected from the Taylor cone. Some works have
pointed out that the solvent evaporation starts rapidly from the jet surface and causes
the jet to become extremely unstable. So, the PVP concentration of jet surface sol
increases sharply and is much larger than that of jet internal sol. If the PVP
concentration of the surface sol is increased to a coagulated critical value at a
relatively earlier stage, in which the jet travels steadily and extends along a single
straight line where the PVP sol is transforms into PVP gel. Namely, the viscous PVP
sol shell begins to freeze and transforms to an elastic skin. But the internal sol is still
viscous flow. Once the elastic behavior of skin overcomes viscous behavior of
internal sol, the columnar flow jet is immediately buckled into a flattened flow jet
(Fig. 7d) with approximately rectangular cross section (Fig. 7e). Subsequently, the
25 Ferrites Obtained by Sol-Gel Method 709

Fig. 8 (a, b) Representative FESEM, (c) TEM, and (d) HRTEM images of SrFe12O19 nanoribbons:
PVP concentrations of 12.3% (Jing et al. 2015)

jet is subjected to an unstable stage with a series of bending instabilities and


anisotropic shrinkage and is finally elongated with or without branching/splitting
and solidified to PVP/SrFe12O19 composite precursor nanoribbons (Fig. 7f).
Figure 7g shows the SEM image of the collected PVP/SrFe12O19 composite precur-
sor nanoribbons for 8.5% PVP concentration. When electrospinning is finished, the
as-spun PVP/SrFe12O19 precursor nanoribbons are annealed at the temperature of
800  C in air, the PVP is degraded completely, and the Sr2+ and Fe3+ ions are
compounded to SrFe12O19. Finally, the SrFe12O19 nanoribbons (Fig. 7h and i) are
constructed by crystalline SrFe12O19 nanoparticles.
It is revealed that in the SEM images (Fig. 8a and b), all the samples present a
novel ribbon-like structure constructed by a large number of inter-connecting
SrFe12O19 nanocrystalline particles. The average widths are estimated to be about
484  15, 812  10, and 1099  18 nm for 8.5%, 10.4%, and 12.3% PVP
concentration, respectively, indicating that the ribbon width broadens with the
increase in PVP concentration in their spinning solutions. Some pores can be
observed obviously in the SrFe12O19 nanoribbons. These pores are caused by the
PVP decomposition and SrFe12O19 nanoparticles crystallization during the heat
treatment. Pores shrink gradually with the increase in PVP concentration from
8.5% to 12.3% and the reduced SrFe12O19 nanoparticles distribute more evenly
and densely, which make the nanoribbon surface more smooth and neat. Each
nanoribbon has been demonstrated to be structurally continuous and almost identical
width that assembled well by abundant single-domain SrFe12O19 nanoparticles
along the long-axis direction. Besides, the PVP concentration in the spinning
solution has a significant influence on the ribbon width, surface smoothness, and
particle size of SrFe12O19 nanoribbons. Considerable high Ms and Hc were observed
for all SrFe12O19 nanoribbons that further increased with the ribbon-width broaden-
ing. The highest Ms of 67.9 emu/g and Hc of 7.31 kOe were concurrently acquired
by SrFe12O19 nanoribbons with the highest ribbon width.

Ferrite Thin Film

Bulk materials are widely used in electroacoustic devices, motors, electricity


generators, etc. On the other hand, thin films of ferrites are widely applicable in
the fabrication of magnetic and magneto-optic devices, such as high-density
710 S. E. Shirsath et al.

recording media because of their excellent chemical stability, low corrosion,


moderate mechanical durability, and considerably high uniaxial magnetic anisot-
ropy. In the thin film form, ferrites were proposed to be the important materials
for nonreciprocal microwave devices because of its high electrical resistivity and
frequency dependent permittivity. Surface morphology of the ferrite thin films is
crucial to the signal-to-noise ratio (SNR) in high density magnetic recording
media. Precise control over morphologies of bulk/thin film materials is an
important aspect from the point of view of their applications and also important
in material science research. It is a result of atom immigration in nonequilibrium
or near equilibrium state (Snelling 1988; Morisako et al. 1997; Zhang et al.
2012a).
In order to develop the ferrite thin films particularly for high-density magnetic
recording media, the films should exhibit high Hc (>3 kOe) to record information
bits with narrow transition length, a high remnant magnetization (Mr), as well as unit
remnant ratio (Mr/Ms ~1) to achieve high SNR. Uniformly distributed magnetic
nanoparticles must have small particle size (<10 nm) to satisfy the requirements for
high storage of magnetic data. The size of particles affect the reduction of magnetic
domain boundaries, which leads to single-domain particles and a thermal instability
of the total spin system (called as superparamagnetism). The room temperature
single-domain diameter is estimated as 100 nm (Kittel formula), whereas it is
10 nm (Neel theory) for the superparamagnetic diameter. That makes it even more
important to reduce the magnetic particle size in thin films without affecting their
magnetic properties for the application in low-noise high-density magnetic recording
media (Pramanik et al. 2005).
Several methods have been employed for the preparation of thin films, such as
pulsed laser deposition (Eason 2007), sputtering (Shirsath et al. 2016), electron beam
evaporation (Li et al. 2016), metallo-organic decomposition (Díaz-Castañón et al.
2014), and sol-gel processing (Leu et al. 2015; Solovyova et al. 2015). Among the
available synthesis methods, sol-gel method that uses aqueous solution of metal salts
has recently emerged as a versatile technique for the preparation of thin films.
Different types of inorganic materials in thin film form with a better homogeneity
in a cost-effective manner can be prepared via sol-gel thin film technique. The major
advantage of such technique is to produce crystallized ferrite particle in thin film
form at lower annealing temperature that have a good control over the particle size
formation and their uniform distribution, which is a main requirement for their
applicability in high-density magnetic recording media. It has the capability for
obtaining narrow grain size distribution in thin films, high solvability of precursors
in final sol, and obtaining suitable magnetic features without using expensive
techniques such as PLD and sputtering.
In general, sol-gel process utilizes metal nitrates as the starting materials; how-
ever, the metal chlorides as starting materials can also be used. It is a known fact that
the silicon (Si) substrates are widely used as the substrate for electronic devices;
growth of ferrite thin films on Si substrates by the sol-gel method did not attract the
attention of the researchers because of the formation of inhomogeneous film distri-
bution on Si substrate from aqueous precursors. The appearance of the
25 Ferrites Obtained by Sol-Gel Method 711

inhomogeneity in the desired thin films might be due to the viscosity of the aqueous
solution that is not suitable for the application of the sol-gel assisted spin-coating
method (Pramanik et al. 2005).

Spinel Ferrites

Sol-gel chemistry combined with polymer templating strategies enables the synthesis
of ferrite thin films in an ordered single phase cubic network of nanodimension pores
as well as tunable spinel structured domain sizes. Polymer templating strategies are
considered to be an efficient route to synthesize nanocrystalline metal oxide frame-
work with desired mesoporous morphologies. The synthesis of these materials is
heavily relying on the solution phase coassembly of inorganic sol-gel precursors
with an organic structure-directing agent, in amphiphilic polymers. The corresponding
thin films can be achieved by the same coassembly methods by using an evaporation-
induced self-assembly process. C. Reitz et al. focused on mesoporous CuFe2O4 spinel
thin films, materials that can be readily synthesized with a 3D honeycomb structure as
well as tunable nanocrystalline magnetic domain sizes (Reitz et al. 2012). In their
work, they incorporated a KLE-type diblock copolymer (here H
[(CH2CH2)0.67(CH2CHCH2CH3)0.33]89-(OCH2CH2)79OH) as the structure directing
agent and hydrated copper (Cu(NO3)23H2O) and ferric nitrate (Fe(NO3)9H2O)
salts as the starting materials. Top view scanning electron microscopy images in
panels (a) and (b) show a cubic network of interconnected pores of ~17 nm in
diameter; the total percentage porosity is ~30% (Fig. 9). It is revealed from these
data that the presence of major structural defects can be ruled out following the
nanocrystalline spinel materials through amorphous material. The pore walls are up
to 17 nm thick, which helps to understand retained nanoscale structure after the
crystallization. TEM images shown in panels (c) and (d) are an evidence of honey-
comb pore network at the hexagonal top surface that also persist throughout the
CuFe2O4 thin films. It is also evidenced from TEM images that the pore walls exhibit
particulate nanodomains. The average magnetic domain size is in good agreement
with the pore wall thickness obtained from SEM images. TEM selected area electron
diffraction (SAED) pattern is shown in the inset in panel (c) that confirms the existence
of typically diffuse Debye–Scherrer rings corresponding to a nanocrystalline but
randomly oriented domain of spinel ferrite material. The crystalline nature is also
confirmed by high-resolution TEM (HRTEM), as can be seen in the inset in panel (d).
It is evidenced from both SEM and TEM images that the spinel ferrite materials are
well defined at the nano/microscale. Grazing incidence small-angle X-ray scattering
(GISAXS) technique was used to collect the quantitative information related to pore
structure of the KLE-templated CuFe2O4 thin films (Fig. 9e–g). The observation
revealed the formation of noncrystalline (amorphous) CuFe2O4 possessing face-
centered cubic (fcc) close-packed structure with preferred (111) crystal orientation.
Large unidirectional lattice contraction is revealed by the elliptical shape of the
GISAXS pattern (Fig. 9e). However, because the CuFe2O4 thin films are engineered
by sol-gel methods that bound firmly to the substrate, this work shows that CuFe2O4
712 S. E. Shirsath et al.

Fig. 9 (a–d) Morphology and nanoscale structure of KLE-templated CuFe2O4 spinel thin films
heated to 650  C. (a, b) Top view SEM images. (c, d) TEM images. The inset in panel (c) shows a
SAED pattern. The inset in panel (d) displays a HRTEM image showing the (101) lattice planes of
tetragonal CuFe2O4. (e–g) Synchrotron-based GISAXS at an angle of incidence β = 0.2 on thin
films heated at 300  C for 12 h (e) and 600  C (f) and 700  C (g) for 10 s, respectively (Reitz et al.
2012)

can be template to synthesize high-quality mesoporous magnetic thin films with a


partially inverted spinel structured domain. CuFe2O4 thin film synthesized by this way
exhibit multiple functionalities that include redox and photoactivity, therefore could
pave the way for prototype device design and fabrication.
A considerable enhancement of magneto-optical effects has been achieved in
CoFe2O4 upon the substitution of Mn3+ (CoFe2xMnxO4). Curie temperature (Tc) of
Mn3+ substituted CoFe2O4 films decrease linearly, resulting in some abnormal
behavior in microstructure and magnetic properties observed in the CoFe2xMnxO4
films (Zhou et al. 2001). To investigate these abnormalities, CoFe2xMnxO4 (where,
x = 0–2.0) spinel ferrite nanocrystalline thin films and nanopowders were synthe-
sized by a modifying Pechini-type sol-gel method (Zhou et al. 2002). Here, stoi-
chiometric amount of Fe(NO3)39H2O (ferric nitrate nanohydride), Co(NO3)26H2O
(cobalt nitrate hexahydride), and Mn(NO3)2 (manganese nitrate) solutions (50 wt%)
were dissolved in water–alcohol mixed solution. Polyethylene glycol and citric acid
were used to synthesize the polymeric precursor solution consisting uniformly
dispersed metal ions. Spin-coating method was used to deposit the CoFe2xMnxO4
thin film on a single-crystalline Si(100) substrate. The speed of the rotator was
maintained at 4000 rpm for 8 s. The as-synthesized thin film was dried in air
followed by annealing in oxygen atmosphere to remove the organic and volatile
species. It has been observed that Mn3+ substituted CoFe2O4 ferrite thin films
possesses an inverse spinel structure like the pure Co ferrite for the doping
25 Ferrites Obtained by Sol-Gel Method 713

concentration x = 0 1.5. Further increase in substitution of Mn3+ where x > 1.5


resulted in migration of Co2+ ions to tetrahedral A-sites, whereas Mn3+ ions prefer to
occupy octahedral-B sites; therefore, the crystal structure of CoFe2xMnxO4
(x > 1.5) films was changed to a normal spinel. Owing to the fact the transformation
of cubic phase, no secondary or impurity phases were observed in these materials.
On the other hand, for CoFe2xMnxO4 over the entire range of Mn3+ substitution
nanocrystalline form with cubic single phase maintained by the modified Pechini-
type synthesis technique. Transformation of cubic to tetragonal (c ! t) phase could
be observed in powder form at x = 1:3 for CoFe2xMnxO4. The possible reason for
the discrepancy in the phase transformation point between CoFe2xMnxO4 powder
and thin film could be related to the single crystalline Si substrate with cubic
structure that keeps cubic CoFe2xMnxO4 more stable. The saturation magnetization
for nanocrystalline powder is higher as compared to their thin film counterpart.
Ni-Cu-Zn ferrites are known to be excellent soft magnetic materials for their
application in high-frequency devices such as multilayer chip inductors (MLCI)
because of its low eddy current losses, high frequency dependent permeability/
dielectric constant, high electrical dc resistivity, and low cost. Semiconductor indus-
tries prefer rapid thermal annealing (RTA) treatment that crystallize the materials in a
short period of time but lower the possibilities to setup interference between the
semiconducting thin film and substrates. Studies revealed that RTA for sol-gel
synthesized materials is very promising to fabricate the device at relatively low
annealing temperature. Feng Liu et al. fabricated Ni0.8xCu0.2ZnxFe2O4 thin films
via sol-gel technique followed by RTA treatment (Liu et al. 2007). They investigated
the effect of varying sintering time, temperature, and chemical composition on Ni-
Cu-Zn ferrite thin film for their structural and magnetic properties. The analytical
grade iron nitrate [Fe(NO3)3], nickel nitrate [Ni(NO3)2], dimethylformamide
[C3H7NO], zinc acetate [ZnC4H6O4], and copper nitrate [Cu(NO3)2] were used as
starting chemical. These chemicals in desired stoichiometric proportion were
dissolved in dimethylformamide solution to obtain Ni0.8xCu0.2ZnxFe2O4. The
mixed solution was stirred for 1 h and then acetic acid (CH3COOH) was introduced
to control the 0.3, 0.4, and 0.5 mol/L concentration of the mixed solution. In the
mean time, polyvinylpyrrolidone (C6H9NO)n surfactant was added to prevent che-
lating of colloidal particles from each other. In order to form a stable sol-gel
precursor, the as-synthesized solution was stirred continuously for 2 h and kept at
room temperature for 36 h. N-type (100) oriented Si substrate of 900–1000 Ω cm
resistivity coated with 0.2 μm SiO2 was used to deposit Ni-Cu-Zn ferrite thin film. To
pyrolyze and exclude the organic substances, the as-synthesized ferrite films were
presintered for 10 min at 120  C and followed by annealing for 30 min at 400  C.
The cycles of the operation of spin coating, drying, and heating were repeated to
reach the required thickness of the films. Then, RTA process for the period of
1–10 min in the temperature range of 400–700  C with an increment of 150  C/s
was carried out in oxygen atmosphere to achieve the complete crystallization of Ni-
Cu-Zn ferrite thin film. As evidenced from the XRD analysis, that sol-gel synthesis
effectively lowers the crystallization temperature of Ni-Cu-Zn ferrite thin film. It has
been observed that the saturation magnetization of Ni0.8xCu0.2ZnxFe2O4 thin film
714 S. E. Shirsath et al.

increased up to 271 emu/cm3 for Zn2+ substitution (x = 0.45) followed by decrease


in magnetization with further increase in x. On the other hand, coercive field attains a
maximum value of 15.62 Oe for Zn2+ substitution of x = 0.4. The experimental
results confirmed that the Ni-Cu-Zn ferrite spin coated thin film of concentration of
0.4 mol/L thoroughly crystallize at 600  C (RTA = 5 min) that exhibit remarkable
soft magnetic properties with low coercivity, Hc = 15.62 Oe, and moderate satura-
tion magnetization of Ms = 228.88 emu/cm3 (Liu et al. 2007).
In sol-gel synthesized thin films, nucleation and growth during sintering process
govern the film microstructure. Therefore, the manipulation of the microstructure
can be obtained by varying the synthesis conditions, viz., chemical precursor
composition, heat treatment parameters, and appropriate substrate selection. Mini-
mizing the magnetic anisotropy is at-most important to alleviate magnetostriction in
transformers and magnetic inductors applications. Hence, S. Seifikar et al. have
grown the randomly oriented NiFe2O4 ferrite thin films on Si substrates by manip-
ulating annealing and pyrolysis conditions that certainly affect the magnetic prop-
erties and microstructure of NiFe2O4 ferrite thin film (Seifikar et al. 2014).
NiFe2O4 solution of 0.5 molarity is synthesized by dissolving nickel acetate
tetrahydrate [Ni(CO2CH3)24H2O] and ferric nitrate nonahydrate [Fe(NO3)39H2O]
in a 2-MOE(CH3OCH2CH2OH) solvent separately. 2-MOE is used primarily due to
its ability to solubilize a variety of starting reagents. NiFe2O4 thin films were
deposited onto (100) oriented Si substrates by using spin-coating technique. The
deposited films are pyrolyzed on a hot plate for 1 min and rapidly thermal annealed
in air atmosphere. Under equilibrium conditions, the chemical reaction is

NiðCO2 CH3 Þ2  4H2 O þ 2FeðNO3 Þ3  9H2 O þ CH3 OCH2 CH2 OH þ ð3x þ 1=2ÞO2
! NiFe2 O4 þ 7CO2 þ 29H2 O þ 6NOx

XRD data confirmed the formation of single phase cubic spinel NiFe2O4 ferrite
thin film at all the pyrolysis temperatures, and no secondary phase is detected. The
as-deposited NiFe2O4 ferrite thin film is randomly oriented with highest intense
(311) peak that match the X-ray diffraction patterns of bulk powder of NiFe2O4
ferrite (ICDD PDF card no. 10-0325). No difference in crystallinity between the
samples is observed as all of the samples have almost identical XRD patterns.
Despite the different pyrolysis temperatures, all films have comparable average
grain shape and size of the grain about 40 nm in diameter. Cross-sectional SEM
images ( figure not shown here) also show no significant differences in the grain
surface morphology, indicating no major microstructural change as a result of
pyrolysis temperature. This confirms that the pyrolysis step does not affect much
the films’ microstructure, pointing toward thermal budget provided in this step is not
sufficient to overcome even locally the nucleation activation energy.
On the other hand, it has been observed from the XRD that with the increased
annealing temperature, the spinel NiFe2O4 peaks became narrow and XRD intensity
increase, that is an indication of grain growth and more complete crystallinity. The
saturation magnetization increased with annealing temperature that subjected to the
increase in grain size. The film annealed at 900  C has the highest Ms, which is
25 Ferrites Obtained by Sol-Gel Method 715

400
annealing T
300

200
Magnetization (kA/m)

100

0 250

Coercivity field (Oe)


200
–100 150
100
–200
50

–300 0
25 30 35 40 45 50 55 60
Grain size (nm)
–400
–10 –5 0 5 10
Magnetic Field (4pkA/m)

Fig. 10 (a) Magnetization hysteresis loop s of Si/NiFe2O4 films annealed at temperatures ranging
from 650  C to 900  C. The inset shows the coercivity (Hc) versus the NiFe2O4 grain size. (b)
Selected area electron diffraction patterns of the Si/NiFe2O4 thin films annealed at 900  C (Seifikar
et al. 2014)

above the reported value for bulk NiFe2O4 (270 emu/cm3) at 300 K. However, the
observed increase in the saturation magnetization with increased grain size is
contrary to the dependence of the magnitude of Ms on the material chemistry rather
than the microstructure. In ferro/ferrimagnetic materials, the microstructure affects
the shape of the M–H curve (coercivity and remnant magnetization) and the strength
of the magnetic field at which the saturation of magnetization is attained (Fig. 10a).
The SAED patterns also confirm the purity of the nickel ferrite and absence of any
secondary phase (Fig. 10b). Therefore, the high Ms in the Si/NiFe2O4 thin films
annealed at 900  C is not associated to the presence of any magnetic impurity or
secondary phases (Seifikar et al. 2014).
Hydrogen is not only highly demanding in space, aeronautics, and automobile
industries as a promising green energy source, but it is also applicable as reducing
agent in a variety of chemical industries. Highly sensitive, economic, and easy to
operate semiconductor based metal oxide (SMO) sensors are in great demand for
hydrogen gas detection. For long-term applications, compact thin film sensing
elements with integrated heater have been considered much more attractive than
discrete Taguchi-type sensing elements. The hydrogen sensing characteristics of
magnesium zinc ferrite (Mg0.5Zn0.5Fe2O4) thin film sensor are reported by manip-
ulating the film thickness, sensor operating temperature, and test gas concentration
(Mukherjee and Majumder 2014).
Precursor sol for magnesium zinc ferrite thin film was synthesized via nitrate salts
of magnesium, zinc, and iron. Citric acid and ethylene glycol were used as chemical
additives to stabilize these precursor sols. The prepared sol is deposited on alumina
substrate and immediately presintered at 400  C for 5 min and then quenched
quickly to room temperature. The number of firing cycles and coating (8, 15,
30, and 45 times) were repeated to increase the Mg0.5Zn0.5Fe2O4 film thickness.
Finally the as-coated films were annealed at 600  C for 2 h in air for complete
716 S. E. Shirsath et al.

Fig. 11 Schematic
architecture of the thin film-
based sensing element
(Mukherjee and Majumder
2014)

crystallization. The schematic of thin film configuration for gas sensing measure-
ment is shown in Fig. 11. Inter-digitated gold electrodes were sputter deposited on
thin film surface to measure the resistance transient upon exposure to air and test gas
environment.
The gas sensing characteristics of the synthesized Mg0.5Zn0.5Fe2O4 films were
characterized by systematically manipulating the film thickness, operating temper-
ature, and test gas concentration of the sensing elements. The response of the sensor
was determined using the following relation:

Response ð%Þ ðSÞ ¼ ðRo  RhÞ=Ro  100 (5)

where (Ro) and (Rh) are the measured value of equilibrium resistance in air and
hydrogen, respectively.
During the detection of test gases, the response and recovery times of these
Mg0.5Zn0.5Fe2O4 thin film sensors were determined as the times required for 63%
change of their resistance from the resistance measured in air [response time (τres)] or
the resistance measured in test gas [recovery time (τrec)]. The resistance transients for
1660 ppm hydrogen sensing using 8, 15, 30, and 45 times coated Mg0.5Zn0.5Fe2O4
films are presented Fig. 12a. The hydrogen “on” and “off” states are marked in the
figure as up (") and down (#) arrows, respectively.
The resistance of the sensor decreases when the hydrogen gas is turned “on,”
where the sensor recovery is achieved by flowing 500 sccm air into the gas sensing
chamber as the hydrogen flow is turned “off.” Minimal resistance drift is observed
when the thin films are switched back and forth between air and hydrogen. The
15 and 30 times coated Mg0.5Zn0.5Fe2O4 films possess better resistance change and
faster response and recovery kinetics as compared to the 8 and 45 times coated films.
For 1660 ppm hydrogen sensing at 250  C, Fig. 12b compares the response (%) of all
the synthesized Mg0.5Zn0.5Fe2O4 films with varying film thickness. The response/
recovery times are determined from the respective response transients. The estimated
response/recovery times for these Mg0.5Zn0.5Fe2O4 films are tabulated in the inset of
Fig. 12b. Note that 15 and 30 times coated films yield better response (%) with rapid
response time. Overall, response is found to be rapid than the respective recovery
processes for all the ferrite thin film samples. The longer recovery time could be due
to the mesoporous morphology of these spin coated Mg0.5Zn0.5Fe2O4 ferrite thin
films (Mukherjee and Majumder 2014).
Addition agent glucose as a kind of organic compound can play a critical role of
fuel in the self-igniting combustion process. Glucose is also a low-cost material and
25 Ferrites Obtained by Sol-Gel Method 717

a 80 8 times 15 times
160
75
140
Resistance (MW)

Resistance (MW)
70
120
65
100
60 80
55 60
50 40
45 20
0 200 400 600 800 1000 1200 0 500 1000 1500 2000

80 90
30 times 45 times
70 80
Resistance (MW)

Resistance (MW)
60
70
50
60
40
50
30
20 40

10 30
0 300 600 900 1200 1500 1800 0 300 600 900 1200 1500 1800
Time (s) Time (s)

b
80 Sample Response Recovery
(thin film) time (s) time (s)
8 times coated ~55 ~60
15 times coated ~32 ~160
60 30 times coated ~28 ~200
Response (%)

45 times coated ~55 ~95

40 8 times coated
15 times coated
30 times coated
45 times coated
20

0
0 400 800 1200 1600
Time (s)

Fig. 12 (a) Resistance transients and (b) response transients of Mg0.5Zn0.5Fe2O4 films for the
detection of 1660 ppm of H2 at ~250  C (Mukherjee and Majumder 2014)
718 S. E. Shirsath et al.

abundantly available. Apart from this, glucose possesses macromolecular structure


characteristic that makes it as a good dispersant and reducing agent (Sun et al. 2012).
J. Sun et al. successfully synthesized polycrystalline and nanometric CoFe2O4 ferrite
thin films by using glucose as an addition agent. Glucose (C6H12O6) of 3.0 wt% was
dissolved in deionized water and then mixed with the stoichiometric iron nitrate [Fe
(NO3)39H2O] and cobalt nitrate [Co(NO3)26H2O] mixture. Here, glucose can
prevent magnetic particles from chelating with each other. Finally, the product is
formed with the following chemical reaction:

CoðNO3 Þ2  6H2 O þ 2FeðNO3 Þ3  9H2 O þ C6 H12 O6 þ 2NH3  H2 O


¼ CoFe2 O4 þ 6CO2 " þ3N2 " þ3NO " þNO2 " þ35H2 O (6)

The above chemical reaction is an autocatalytic redox reaction, where the nitrate
ions of Co and Fe act as an oxidizer, while the carbonyl from glucose plays the
critical role of reducing agent.

Hexaferrites

Barium ferrite hexaferrite (BaFe12O19) in thin film form has been considered as one
of the primary candidates in the fabrication of high-density magnetic recording
media because of its high coercivity, moderate saturation magnetization, high
remanant ratio, high degree of crystalline anisotropy, and remarkable chemical
stability. Accordingly, the influence of Ba/Fe ratio, calcination temperature and
basic agent on the structure phase formation, crystallite size, and morphology of
thin films have been extensively investigated (Salemizadeh and Seyyed Ebrahimi
2009). The sols for the BaFe12O19 were synthesized by dissolving iron nitrate [Fe
(NO3)39H2O], citric acid (C6H8O7), and barium nitrate [Ba(NO3)2] in ethylene
glycol [C2H4(OH)2] at continuous stirring at ~80  C. In these solutions, the ratio
of Ba:Fe varies from 8 to 10. The pH of mixed solutions was also kept at 7.0 by
dropwise inclusion of trimethylamine (50% solution) C3H9N and ammonia (25%
solution) NH3H2O. The sols prepared in this way were refluxed for 24 h at 353  C
and then spin coated on (100) oriented Si single crystal substrate that has been
cleaned several times by acetone and distilled water. The spin speed was kept at
2000 rpm where the spin time was 20 s. The samples prepared by this way are
pre-sintered for 24 h at 120  C for 24 h and finally annealed in air for 1 h at three
different temperatures 700  C, 800  C, and 900  C. In this study, it has been
observed that the single-phase BaFe12O19 can be successfully achieved at 700  C
by using C3H9N as a pH-controlling agent. This effect could be probably obtained
due to methyl group (CH3) in C3H9N that improves the reactivity of material in the
sol. The procedures of varying the Ba/Fe ratio and basic reducing agent resulted in
minimizing the crystallization temperature from 900  C to 700  C of single-phase
BaFe12O19 thin films form. BaFe12O19 with nanodisks shape can be clearly observed
in Fig. 13a–c, where the nanodisk size became much smaller with decreasing the
processing temperature.
25 Ferrites Obtained by Sol-Gel Method 719

Fig. 13 SEM photographs of the single-phase Ba-hexaferrite thin films: (a) calcined at 900  C
using ammonia as basic agent with Fe/Ba = 10, (b) calcined at 800  C using ammonia as basic
agent with Fe/Ba = 8, (c) calcined at 700  C using trimethylamine as basic agent with Fe/Ba = 8
(Salemizadeh and Seyyed Ebrahimi 2009)

Ferrite thin film shows distinct properties as compared to their nanoparticle


counterpart. Figure 14 shows the XRD patterns of SrFe12O19 (strontium ferrite)
nanoparticles and thin film synthesized by sol-gel route (Ghasemi 2014). It has been
observed from the XRD patterns of SrFe12O19 nanoparticles that the (107) and (114)
are the two dominant X-ray diffraction peaks (Miller indices) of strontium ferrite
nanoparticles. On the other hand, XRD of thin film confirms the c-axis oriented
growth of hexagonal SrFe12O19 film. Diffusion distances are fairly long for the
sample prepared with conventional high temperature ceramic methods as compared
to the sample prepared with the sol-gel process, where mixing of cations takes place
at the atomic level, that results in reduced diffusion paths for various cations. In case
of SrFe12O19 ferrite powder nanoparticles, some secondary phase of hematite
appeared in the structure. Whereas in case of SrFe12O19 ferrite thin film, there is
no significant sign of any secondary phase. Parameters that affect the single phase
formation of magneto-plumbite hexagonal strontium ferrite phase are the sintering
temperature, pH of the solution, and molar ratio of Fe:Sr. In fact, the molar ratio of
Fe:Sr plays a very crucial role in the formation of the magneto-plumbite SrFe12O19
hexagonal phase in the sol-gel process. Further, there is large difference in the
hysteresis loop of sol-gel-prepared SrFe12O19 nanoparticles and thin film (Fig. 14).
SEM image of SrFe12O19 ferrite nanoparticles (Fig. 15a) shows that the morphol-
ogy of nanoparticles is almost semispherical and the distribution size is narrow. The
mean particle size for nanoparticles is about 45 nm. It can be noticed from the SEM
of thin film (Fig. 15b) that the estimated grain size is ~110 nm with elongated shape.
A densely packed microstructure obtained at annealing temperatures represents the
high uniformity of these ferrite thin films (Fig. 15c). Film generally has good
microstructure with few defects such as pinholes or microcracks.
The orientation of the c-axis plane of the Sr-hexaferrite thin film is random,
giving rise to almost isotropic magnetic anisotropies (Ishikawa et al. 1996). An et al.
(2002) reported that the textured barium hexaferrite films are synthesized by sol-gel
process. However, the magnetic anisotropies are still isotropic. They proposed that
to obtain anisotropic magnetic properties, the randomly oriented grains must be
720 S. E. Shirsath et al.

Fig. 14 XRD patterns and VSM plots (perpendicular) of sol-gel-synthesized SrFe12O19 nano-
particles and thin film (Ghasemi 2014)

Fig. 15 (a, b) SEM of powder nanoparticles and thin film, respectively. (c) AFM of thin film of
SrFe12O19 (Ghasemi 2014)

reduced to obtain the anisotropic magnetic anisotropy. To achieve these goals, Tang
et al. (2010) fabricated barium hexaferrite (BaM) thin films via sol-gel synthesis by
spin-coating route. Polyvinylpyrrolidone (PVP, 2 wt%) was added to the solution
that homogenizes the nitrate, enables desired anisotropic growth of the BaM thin
film, and also improves the nucleation of crystallites along the c-axis perpendicular
to the film plane. Films prepared by this route are highly anisotropic both in the
texture and the magnetic anisotropy as evidenced by VSM and XRD study. XRD
pattern indicates that the BaM films have been epitaxially grown on the Al2O3
sapphire (001) plane where [100] direction of BaM parallel to (110) of Al2O3 and
the parallel c-axes. The coercive fields for the sample annealed for 2 h at 1000  C are
87.5 and 230.8 kA/m for in-plane and perpendicular direction, respectively. The
coercive field and the saturation magnetization increased with the increased sintering
time, and the maximum perpendicular Hc is ~ 320.6.3 kA/m for the BaM samples
and were sintered for 3 h.
This group further extended their study on barium ferrite thin film by varying
growth conditions in order to examine their effect on the material structural and
25 Ferrites Obtained by Sol-Gel Method 721

magnetic properties (Zhang et al. 2012b). Barium hexa-ferrite thin films were spin
coated on sapphire (001) substrate. Ethylenediaminetetraacetic acid (EDTA),
C10H16N2O8, was used to chelate the Ba and Fe metal ions. The conditions to obtain
highly epitaxial BaM thin films were expanded by using EDTA, where the pH of the
solutions were changed from 3 to 9 and the molar ratio of metal ions to EDTA varied
from 1.5 to 1.0. It is observed that under all these optimized synthesis conditions,
faceted hexagonal grains of BaM can be obtained. The amount of EDTA influenced
the nucleation and growth of grains for a fixed pH of 7. Dotty small crystallites in the
films are observed at a molar ratio of EDTA to the metallic ions less than 1.0. The pH
values could be varied from 3 to 9 by adding ammonia (NH3) for the EDTA
ratio  1.0, but the morphologies of the BaM thin films were not noticeably affected.
It has been concluded that at a lower value of EDTA, cations are not being highly
chelated. The cation distribution of metallic ions was not uniform that resulted in
higher nucleation density. The grains could not grow-up for the lower concentration
of EDTA and resulted in the formation of dotted grains rather than a continuous
smooth BaM thin film.
Recently, gas sensing technology is one of the most typical application in
intelligent systems. It became even more significant for future development with a
constantly increasing number of applications in environmental monitoring, indus-
trial production (e.g., methane and detection in mines), pollution/air quality control,
healthcare, automobiles, medical devices (such as electronic noses simulating the
human olfactory system), hydrogen economy, and technical processes control.
Among all available sensor technologies, the metal oxide semiconductor (such as
ferrite based) gas sensors have received considerable attention because of their
advantages of robust design, long lasting, small size, lightweight, high material
sensitivity, quick response times, ease of fabrication, and low cost. It is highly
recommended that these ferrites based metal oxide sensor possess high surface
area to absorb most of the target analyte on the surface for better measurable
response sensitivity. Among the known perovskite metal oxides, Fe3+-substituted
strontium titanate (SrTi1xFexO3δ) ferrite materials (STFx in short) have been
reported as a promising candidate as semiconductor-type oxygen sensors. Chow
et al. (2013) reported the synthesis and fabrication of STFx thin film sensors. It is
very difficult to control the hydrolysis rate of the multicomponent alkoxide such
as STFx. Therefore, sol-gel process was modified by using metal salts such as
nitrates and acetates instead of the conventional sol-gel method, which usually
employs metal alkoxides. Using a magnetic stirrer, the titanium butoxide [Ti
(OCH2CH2CH2CH3)4] was first thoroughly mixed with acetylacetone
[CH3COCH2COCH3] at 60  C with a mole ratio of 1:4. At the same time, iron
nitrate was dissolved in 2-methoxyethanol and acetylacetone at 60  C with a mole
ratio of 1:20:6. Next, the Ti4+ and Fe3+ precursors were thoroughly mixed together
and homogenized under continuous stirring for 30 min. In this process, the role of
chelating agent is played by acetylacetone that can prevent the forming of metal
complexes of the metal ions. On the other hand, the strontium acetate was dissolved
in deionized water and glacial acetic acid (CH3COOH) with a mole ratio of 1:15:90.
Deionized water was added to dissolve the acetate powders, which normally have
722 S. E. Shirsath et al.

Fig. 16 Gas sensing response to 20% oxygen for SrTi1xFexO3δ sol-gel thin films sensors: (a)
STF20 (x = 0.2) and STF40 (x = 0.4) (p-type response); (b) STF60 and STF80 (n-type response);
(c) calculated electrical response values at different operating temperatures; (d) effect of oxygen
concentrations for STF40 and STF80 sensors operated at 300  C (Chow et al. 2013)

lower solubility in water and organic solvents than other metal salts. Finally, the
dissolved Sr precursor solution was slowly added into the mixed Ti and Fe precursor
solution under continuous stirring. The molarity of the final sol was adjusted using
2-methoxyethanol. The as-synthesized solution was kept overnight for 1 day to
allow for sufficient hydrolysis before spin-coating deposition.
The XRD study of the STFx films revealed the formation of single phase cubic
perovskite structure, and the crystallinity of the films was decreased with the increase
in iron (Fe3+) substitution. TEM observation indicated that the STFx films with high
Fe substitution were not crystallized properly and had large proportion of
amorphous-like structure. Gas sensing response of all the prepared films is shown
in Fig. 16. The SrTi1xFexO3δ thin film sensors with x  0.4 showed a normal
p-type sensing response, whereas sensors with x  0.5 exhibited anomalous n-type
sensing response. The switch occurred at STF50 thin film where both p- and n-type
sensing response was noted. Additionally, these n- and p-type sensing responses
were observed for reducing gases and for high oxygen concentration range
(1–100%). The sensing response was co-related to the dominant carriers at the
surface of metal oxide. In this case, the anomalous n-type sensing response was
attributed to the amorphous.
25 Ferrites Obtained by Sol-Gel Method 723

Similarly, the oxygen sensitivity of perovskite SrFeO3δ thin films was examined
by C. Liang et al. (2005). SrFeO3δ films were successfully synthesized on the
alumina substrate via a sol-gel citrate technique at 900  C. SrFeO3δ thin films
behave like p-type semiconducting metal oxide in the temperature range of
377–577  C. The best oxygen sensitivity is observed at 377  C, where the response
time is 2 min as exposed to a change from nitrogen to 0.466% oxygen at 377  C.
The grain size reduction and the valence fluctuation in Fe2+/3+ ions induced by the
substitution of lanthanum (La3+) considerably improve the microwave absorption
properties of strontium ferrite. Y.Q. Li et al. investigated the feasibility of producing
rare earth lanthanum (La3+) substituted strontium ferrite, SrLaxFe12xO19, thin films
using sol-gel and self-propagating high-temperature synthesis to investigate the
co-relation between crystal structure quality, microwave absorption, and magnetic
properties (Li et al. 2011). The ratio of Sr:Fe was 1:9, the ratio of Sr:La varied from
1:1 to 0, and the molar ratios of glycol to citric acid and citric acid to metallic ions
were 2:1 and 2.5:1, respectively. About 7 mL glycol was added to the homogeneous
transparent solution without adjusting pH value after the above solution was stirred
for a few minutes. Citric acid and glycol were used as the chelating agents. The sol
was completely formed after 12 h continuous stirring followed by 30 min ultrasonic
scattering to improve the chelating. Subsequently, the sol was cooled to room
temperature in the air atmosphere. The possible complex chemical reaction in the
solution is shown in Fig. 17.
SiO2 substrate (10  10  2 mm3) was dipped into the SrLaxFe12xO19 sol
(viscosity of the sol is in the range of 4–12 mPa s) and pulled out slowly at speed
~1.5 mm/s, termed as-deposited substrate and cleaned by using acetone and hydro-
chloric acid in the ultrasonic cleaner. After drying, the SiO2 slide was coated with a
thin film layer of the gel. The sample was subjected to pre-annealing in air for 30 min
at 400  C by keeping annealing rates of 7  C/min and then allowed to cool down to
room temperature and process was repeated for 15 times. The multilayered
SrLaxFe12xO19 thin film synthesized in this process was finally annealed for 2 h
at 900  C. It has been observed from the XRD and EDS results that the phase
structure of SrLaxFe12xO19 ferrite thin films is M-type hexagonal where rare-earth
La incorporated in the crystal lattice with the forming secondary phase. The grains
observed on the surface of SrLaxFe12xO19 ferrite film are more densely deposited
and smaller in diameter as compared to pure SrFe12O19, where the grain size is in the
range of ~40–70 nm. Atomic force microscopy (AFM) images confirm that the film
surface is quite smooth and grains are spherical in shape with their diameter in ranges
from 100 to 200 nm. Magnetic property measurements revealed that the highest
coercivity of SrLaxFe12xO19 ferrite films is observed to be 5986 Oe for the x = 0.2.
The higher coercivity and higher area under the hysteresis curve of synthesized
SrLaxFe12xO19 ferrite thin film pointing toward its applicability for electromagnetic
absorption. Further, the dielectric loss angle tangent (tanδ) of SrFe12O19 ferrite films
is considerably decreased with La3+ substitution. SrLa0.2Fe11.8O19 shows the broader
microwave absorption frequency and higher tanδ >0.1 in the frequency range of 9 to
10.5 GHz (Li et al. 2011).
724 S. E. Shirsath et al.

Fig. 17 The complex CH2COOH CH2COO


reaction process in the
solution

C(OH)COOH + Sr2+ → C(OH)COOH Sr+ 2H+

CH2COOH CH2COO

CH2COOH

3C(OH)COOH + 2Fe3+ → Fe2(C6H6O7)3 + 6H+

CH2COOH

CH2COOH

3C(OH)COOH + 2La3+ → La2(C6H6O7)3 + 6H+

CH2COOH

The crystal phase formation of gel is generally being controlled by adding base
catalyst. It has been observed that this phase formation sol-gel synthesized pre-
cursors can be enhanced using vacuum extraction (VE) and is a promising alterna-
tive for base catalyst. Two different procedures need to be followed in precursor
solution synthesis. The solution can be in vacuum in the first stage, and then
immersed in a water bath at around 80  C. The nitric acid and water has to be
extracted from the sol-gel precursor system in order to control the viscosity of the
precursor. The low-viscosity liquid precursors are preferred for spin coating to have
a better control over the thickness of the film, and therefore the extraction process
must be monitored carefully. In one of the other route, the nitrate salts can be directly
dissolved into the solvent. Liu and Wu (2001) applied the vacuum extraction process
for precursor preparation of Ba/Fe = 4:11 ratio which assists the proper stoichio-
metric BaM ferrite phase formation. The pH of the precursor increased with the
extraction of nitric acid by VE treatment that resulted in iron gel formation and
barium ion adsorption by the gel. The resultant precursor thus has better chemical
homogeneity and also prevents the formation of hematite (Fe2O3) and barium oxide
(BaO) phases in BaM ferrite. It is to be noted that BaM ferrite can be obtained by
optimizing Fe/Ba ratio without utilizing the VE treatment. Addition of barium nitrate
in the starting precursor may increase the pH of solution that further induced the
formation of iron gel. BaM thin films with Ba/Fe = 28:2 ratio produce stochiometric
single-phase BaM ferrite with c-axis out-of-plane orientation and thus having
better magnetic anisotropy as compared to BaM films of Ba/Fe = 4:11 ratio (Liu
and Wu 2001).
25 Ferrites Obtained by Sol-Gel Method 725

Bismuth Ferrite

As a multiferroic material, bismuth ferrite (BFO) is perhaps the only material which
exhibits ferroelectric and G-type antiferromagnetic ordering temperature at
Tc = 1103 K and TN = 643 K, respectively (Catalan and Scott 2009). In recent
years, it has been widely studied due to fascinating fundamental physical properties
and potential applications in the electronic memory storage, spintronic devices, and
integrated microelectronic devices. However, pure BFO thin film still shows high
leakage currents, unsaturation P–E hysteresis loops, and weak ferromagnetic prop-
erties due to the existence of oxygen vacancies and antiferromagnetic structures
(Ederer and Spaldin 2005).
Sr2+-substituted Bi0.89xSm0.11SrxFe0.94(Mn0.04Cr0.02)O3 (hereafter termed as
BSSrxFMC) thin films were prepared on the FTO (Fluorine-doped Tin Oxide)/
glass substrates by utilizing the sol-gel method. Nitrates of the respective metal
ions were mixed together with excess 5 mol% of Bi to compensate the evaporation of
bismuth during the sintering process. Then 2-methoxyethanol which served as a
solvent was added into the raw materials and stirred continuously for 1 h. Then acetic
anhydride (C4H6O3) was added to dehydrate and control the pH value of the solution
with continuous stirring for 5 h (here the volume ratio of acetic anhydride and
2-methoxyethanol was kept as 1:3) that produce a stable precursor solution. The
concentration of the solution was kept as 0.3 mol/L. The entire process was carried
out at room temperature in an ambient atmosphere. The as-obtained precursor
solution was then deposited on the FTO/glass substrates for 15 s at 4000 rpm by
spin-coating technique. The as-deposited BSSrxFMC wet films were dried for 5 min
at 240  C to remove any volatile materials and finally annealed for 10 min at 550  C
in air atmosphere for complete crystallization. The results show that the BSSrxFMC
thin film with x = 0.04 exhibited the best multiferroic properties. The ferroelectric
and ferromagnetic properties of the BSSrxFMC thin films have successfully been
enhanced by sol-gel assisted spin-coating method and rare-earth ion substitution (Liu
et al. 2014).
Transparent films of BiFeO3 can also be deposited on quartz and glass substrates
by spin- and dip-coating technique using the synthesized sol precursor (Das et al.
2010), where dip-coating process produced ~2 μm thick films and spin-coating
process resulted in ~200 nm thin film. Here, the sol was synthesized by
wet-chemical approach using 0.004 mol of both ferric nitrate [Fe(NO3)9H2O] and
bismuth nitrate [Bi(NO3)35H2O]. First, an aqueous solution of ferric nitrate salt was
treated with ammonium hydroxide (25%) to form ferric hydroxide solution from
where a clear solution containing Fe3+ ions were obtained. Double-distilled water
under centrifugation at 4000 rpm was utilized to wash the precipitate followed by its
mixing with 20 mL ethanol. Acetic acid (95%) of 4 mL was then added into solution
under continuous stirring and heat treated for 30 min at 75  C. Hence, clear Fe3+-
containing solution was obtained. Five drops (~0.00419 mol) of diethanolamine
(DEA) were added in this solution with continuous stirring. Then, a necessary
amount of bismuth nitrate was dissolved in 20 mL of ethylene glycol monoethyl
ether (2-ethoxyethanol) solvent. Acetic acid (95%) of 0.5 mL was also added into the
726 S. E. Shirsath et al.

solution and continuously stirred for 2 h until it become transparent. Five drops
(0.00419 mol) of DEA were again added to the solution and stirred for 30 min. It was
then gently mixed into the clear Fe3+-based solution under continuous stirring at
room temperature. The solution immediately turned out to be a clear solution that
was further stirred for 12 h. The films were deposited by spin-coating (4000 rpm for
20 s) and dip-coating (speed 6 cm/min) process on quartz, glass, and indium-doped
tin oxide (ITO) substrates. As expected the film thickness of the dip-coated BiFeO3
depends on the viscosity of the solution, speed of the dipping, and number of
coatings. Whereas the thickness of the spin-coated BiFeO3 films rely on viscosity
of the solution, speed (rpm) of the spin coater platform, and number of coatings.
Ethanol and/or 2-ethoxyethanol was added in the final clear solution to control the
2-ethoxyethanol to ethanol total volume ratio as 1:2, 1:1, or 2:1. The molarities of the
final solution for both Bi and Fe remain at 0.004. The images of the BiFeO3 films
synthesized in this way are presented in Fig. 18.
Annealing temperature effect on the structural, ferroelectric, optical, and pho-
tovoltaic properties of sol-gel synthesized BiFeO3 thin films has been investigated
by Lin et al. (2013). In the first step, ferric nitrate [Fe(NO3)39H2O] was mixed in
the acetic acid (CH3COOH) and 2-methoxyethanol (CH3OCH2OH) and (volume
ratio 1:4) solution to form Fe precursor solution. The solution was continuously
stirred for 30 min at 50  C to completely dissolve the ferric nitrate. Then,
acetylacetone [CH3COCH2–COCH3] was added to control the hydrolysis rate of
ferric nitrate. Dehydration of ferric nitrate was examined by adding acetic anhy-
dride [(CH3CO)2O] to produce a clear and homogeneous precursor solution. Bi
precursor solution was obtained through a similar process. In the second step, Fe
precursor solution was mixed with Bi precursor solution maintaining a molar ratio
of Fe:Bi mixture to 1:1.05, followed by continuous stirring for 30 min at 25  C.
Ethanolamine [HOCH2CHNH2] was then added to control the pH and viscosity of
the precursor solution. Necessary amount of the mixed solvent was added to the
solution to keep the concentration of the precursor at 0.2 mol/L. The resultant
precursor solution was transparent as well as homogeneous. The as-prepared
precursor solution kept was for aging at 25  C for 24 h, and then the thin films
were deposited on the quartz glass and Pt(100 nm)/Ti(30 nm)/SiO2(500 nm)/Si
(500 μm) substrates by spin-coating method.
Thermal analysis of BFO gel is depicted in Fig. 19a. It is observed that one
endothermic reaction occurs at 80  C, leading to a weight loss around 16.06% due to
removal of organic solvents and dehydration of residual water. Further, there are two
dominant exothermic peaks observed at 220  C and 275  C, respectively. The
significant weight loss around 47.47% (34.89% + 12.58%) is observed in the
range of 180  C to 400  C corresponding to pyrolysis of organic residues. Apart
from this, there is a small endothermic peak around 593  C, but the weight is
unchanged implying the crystallization of BFO thin films. It has been observed
from Fig. 19b that the fundamental absorption edge for these samples is around
1.62 eV (corresponding to 766 nm), confirming the utility of BFO thin film to absorb
the visible light. The absorption coefficient (α) of the 650  C annealed sample is
comparatively lower than 550  C and 600  C annealed samples and can be related to
25 Ferrites Obtained by Sol-Gel Method 727

Fig. 18 (a, b) Fe-based and Bi-based solutions used to prepare Bi–Fe-based solution as shown in
(d). (c) The thick furnace-cooled transparent BiFeO3 film. (e, f) The cross section of the dip-coated
and spin-coated films, respectively, which indicate that the dip-coated films (6 cm/mm) are thicker
than the spin-coated (4000 rpm/20 s) films. (g) The TEM image of the nanoparticles obtained from
final BiFeO3 solution by putting a drop of it on Cu grid and annealed it at 400  C for 10 min, while
(h) shows the TEM image of the powder BiFeO3 sample obtained after drying solution as shown in
(d) and annealed at 400  C for 30 min (Das et al. 2010)

film quality or impurity phases resulting from the higher annealing temperature. The
optical band gap of BFO thin films was determined using Tauc’s law:

ðαhνÞn ¼ Aðhν  EgÞ

where A is a constant (indicating how steeply the absorption increase with energy), α
is the absorption coefficient, hν is the photon energy, n represents the type of optical
transition between valence and conduction bands, and Eg is the Tauc gap or optical
band gap. The observed band gap of the BFO thin film lies in the range of 2.306 and
2.453 eV, pointing toward the visible light absorption efficiency of BFO thin films
(Fig. 19c). The band gap of BFO thin films annealed at 550  C is higher as compared
to band gap obtained for 600  C and 650  C thin film and could be related to
728 S. E. Shirsath et al.

Fig. 19 (a) Thermal analysis (DTA–TG) curves of BFO gels, (b) transmission and absorption
spectra, and (c) dependence of (αhν)2 on hν of BFO thin films annealed at three different
temperatures (Lin et al. 2013)

crystallinity of the sample. In terms of photovoltaic characteristic, it is found that the


short circuit photocurrent density decreased with the increase in annealing temper-
ature. The sample annealed at 550  C possess the higher open-circuit photovoltage
(Voc) as compared to BFO thin film annealed at 600  C and 650  C. This higher
photovoltaic output for 550  C annealed BFO thin film could be related to higher
internal depolarization field due to the larger remnant polarization. Further, the
maximum power conversion efficiency (ηmax) is observed for 550–650  C annealed
samples and occurs at about half of the open-circuit photovoltage. The ηmax of BFO
thin films annealed at 650  C, 600  C, and 550  C is 0.00366%, 0.00203%, and
0.01617%, respectively. The power conversion efficiency is related to the departure
and mobility of charge carriers guided by the polarization. The higher remnant
polarization of the 550  C annealed samples provides the higher electric field as a
driving force to depart the light electron hole carriers and resulted in to higher power
conversion efficiency. However, it is to be noted that BFO thin films annealed at
600  C possess higher remnant polarization and lower power conversion efficiency
as compared to BiFeO3 annealed at 650  C (Lin et al. 2013). Overall, BFO thin films
annealed at 550  C possess single phase perovskite structure having narrow band
gap, larger remnant polarization, comparatively lower leakage current and coercive
electric field (Ec), larger power conversion efficiency, and photovoltaic output.
25 Ferrites Obtained by Sol-Gel Method 729

Cai et al. also observed photovoltaic improvement based on enhancement of


ferroelectric property and optical band gap in Ti4+ substituted BiFeO3 ferrite thin
films prepared via sol-gel spin-coating method (Cai et al. 2014). This study revealed
that the substitution of Ti4+ prevents the grain growth and enhances the film surface
smoothness of BiFeO3 thin films. Substitution of Ti4+ ions improves the visible
ultraviolet light absorption characteristics and decreases the band gap of BiFeO3 thin
films. Proper amount of Ti4+ ions substitution considerably reduced the leakage
current density of BiFeO3 thin films. Ohmic conduction mechanism possibly lead to
the leakage current density in pure and Ti substituted BiFeO3 thin film at low electric
applied field, whereas space-charge-limited conduction mechanism takes over for
Ti4+-substituted BiFeO3 thin films at higher electric field region. The remnant
polarization (Pr) and coercive electric field, Ec, of Ti4+-doped BiFeO3 thin films is
higher than that of pure BiFeO3 thin films where the coercive electric field of BiFeO3
thin films with 2 at.% Ti4+ is minimum. Remnant polarization and lattice distortion
governed short-circuit photocurrent density of BiFeO3 thin films decreases with Ti4
+
-doped first for 2 at.% Ti4+ and then increases for 4 at.% Ti4+, and vice versa for
circuit photovoltage. The combined effect of increased remnant polarization and
decreased band gap leads to increase in the power conversion efficiency of sol-gel
synthesized Ti4+-doped BiFeO3 thin films (Cai et al. 2014).
Pure BiFeO3, BiFeO3–NiFe2O4, and Sm-doped Bi0.89Sm0.11FeO3–NiFe2O4 thin
films deposited on FTO/glass (SnO2:F) substrates via a sol-gel method modify the
leakage current characteristics and ferroelectric and ferromagnetic properties of the
BiFeO3–NiFe2O4 thin films (Liu et al. 2015). The NiFe2O4 precursor solution was
deposited by spin coating onto the ultrasonically washed FTO/glass substrate. The
as-synthesized wet films were pre-sintered for 5 min at 240  C and then annealed for
10 min at 600  C air atmosphere for complete crystallization and the whole process
was repeated for four times. Then, the Bi0.89Sm0.11FeO3 precursor solution was
deposited by spin coating on the NiFe2O4 film at 4000 rpm for 15 s. Finally, the
samples were pre-sintered for 5 min at 240  C and then 10 min at 550  C.
The thickness of the NiFe2O4 and BiFeO3 layer of the BiFeO3–NiFe2O4 thin film
obtained from cross-sectional SEM is approximately 160 and 640 nm, respectively
(Fig. 20a). Fig. 20b shows the room temperature variation of leakage current density
(J) versus electric field (E) of all the synthesized thin films. The leakage current
density of the BiFeO3–NiFe2O4 thin film is considerably lower than the pure BiFeO3
over the entire region of applied electric field. The measured leakage current density
of the BFO, BFO–NFO, and BSFO–NFO thin films at 150 kV/cm is 1.25  10–3,
1.41  10–4, and 4.55  10–6 A/cm2, respectively. Moreover, the breakdown
electric field of the BiFeO3–NiFe2O4 thin film is 280 kV/cm and is higher as
compared to pure BFO thin film (160 kV/cm) (Liu et al. 2015).
It is noted that the work on sol-gel synthesized BiFeO3 also possess high Ms
values, where the presence of some of the Fe ions in the 2+ valance state
and accompanied by oxygen vacancies rather than attributed to strain engineering
in the BiFeO3 thin film. BiFeO3 thin film synthesized with anisotropic strain
states using templated porous nanostructures shows high magnetization as well as
large modifications in magnetization while applying an external electric field
730 S. E. Shirsath et al.

Fig. 20 (a) SEM cross-sectional images of BSFO–NFO thin film and (b) leakage current densities
of the BFO, BFO–NFO, and BSFO–NFO thin films (Liu et al. 2015)

(Quickel et al. 2015). Where porous BiFeO3 films were synthesized using
coassembly of amphiphilic diblock copolymer templates with sol-gel-type inorganic
precursors. Sintering process was carried out to decompose the polymer and crys-
tallize the BiFeO3, forming a crystalline film with a homogeneous, continuous
25 Ferrites Obtained by Sol-Gel Method 731

nanoporous architecture. As control samples, nontemplated conventional sol-gel


thin films were also prepared for the sake comparison with the templated porous
films. H[(CH2CH2)0.67(CH2(CH)CH2CH3)0.33]89-[OCH2CH2]79OH (referred to as
KLE-22) was used as the organic template for porous films. A scintillation vial
was charged with Bi(NO3)35H2O (0.371 g), Fe(NO3)39H2O (0.310 g), and 2 mL
anhydrous 2-methoxyethanol and allowed to stir for several hours to prepare a
mesoporous BiFeO3. In a separate vial, the diblock copolymer KLE-22 (0.085 g)
was merged into 1 mL of ethanol and heat treated at 50  C till dissolved completely.
The two solutions were then mixed thoroughly and stirred for 2 h and then filtered
for deposition. Films were synthesized by dip coating technique in a humidity-
controlled environment (25–25%) by varying withdrawal rates (between 1 and
2 cm/s) to obtain the desired thickness of the film. The as-deposited films were
sintered in air atmosphere from 50  C to 180  C for 3 h and can soak for 12 h.
Finally, the films were sintered in air at 475  C with temperature increment at 4  C/
min, soaked for 1 min and then immediately removed to obtain crystallization of the
mesoporous BiFeO3 films.
The porous and amorphous BiFeO3 frameworks indicate that the materials
possess homogeneous porosity over the entire film. Amorphous BiFeO3 films
show scattering maxima owing with a disordered cubic pore network as indicated
by the two-dimensional small angle X-ray scattering (2D–SAXS). These pores are
anisotropic, with significant perpendicular pore shrinkage during sintering that is not
balanced by in-plane contraction due to the films firmly attached to the Si substrate.
Some kind of rearrangement/disordering in the mesoporous structure has been
observed while sintering the film to crystallize the framework; however, the porosity
maintains homogeneity. The partial rearrangement resulted in reduced out-of-plane
scattering in 2D–SAXS measurements and could be related to thickness of film;
however, in-plane scattering was still observed. Nanoscale homogeneity and SAXS
scattering could be lost while sintering templated films for longer period of time to
obtain complete crystallization where the cracks were formed in the film to release
in-plane strain. Similar cracks can be formed if the film thickness is too high, as the
strain energy within the film overtakes the interfacial adhesion. The mesoporous and
dense BiFeO3 film possess single phase perovskite structure and is evidenced by
high-angle X-ray diffraction measurements.
Such novel sol-gel assisted approach indicates that porosity can drive the strain
mechanism in the mesoporous and dense thin film. Relatively high surface area of
nanoporous materials also provide a pathway to design mixed composite materials
where the pores can host to multiple materials. In this way, a mesoporous BiFeO3
films were deposited on Pt/Si wafers and were electroplated with a ferromagnetic
nickel phase (Quickel et al. 2015). Easy synthesis of nanostructured BiFeO3, where
the ferrimagnetism in thin film can be electrically switched, could be applicable in a
variety of technological devices. Porous materials therefore offer great potential for
strain engineering in thin film by employing wet chemical methods without the use
for expensive epitaxial substrates. Porosity driven strain can also be reproducible and
tuned by manipulating temperature and without the precision of epitaxial films.
Electrically induced variation in both strain and saturation magnetization in porous
732 S. E. Shirsath et al.

BiFeO3 thin film demonstrates novel magneto-electric coupling that could not be
observed in epitaxially engineered strained thin film (Quickel et al. 2015). These
promising features of sol-gel synthesized mesoporous thin film point toward the
utilization of strain to convert antiferromagnetic BiFeO3 into a ferrimagnetic mag-
netoelectric material.

References
Albuquerque AS, Tolentinoa MVC, Ardisson JD, Moura FCC, de Mendonc R, Macedo WAA.
Nanostructured ferrites: structural analysis and catalytic activity. Ceram Int. 2012;38:
2225–31.
An SY, Lee SW, Shim IB, Kim CS. Growth of nanocrystalline barium ferrite thin films by sol–gel
method. Phys Status Solidi A. 2002;189:893–6.
Banerjee S, Kumar A, Sujatha DP. Preparation of nanoparticles of oxides by the citrate-nitrate
process. J Therm Anal Calorim. 2011;104:859–67.
Bhattacharjee K, Ghosh CK, Mitra MK, Das GC, Mukherjee S, Chattopadhyay KK. Novel
synthesis of NixZn1xFe2O4 (0  x  1) nanoparticles and their dielectric properties. J Nanopart
Res. 2011;13:739–50.
Cai W, Fu C, Gao R, Jiang W, Deng X, Chen G. Photovoltaic enhancement based on improvement
of ferroelectric property and band gap in Ti-doped bismuth ferrite thin films. J Alloys Compd.
2014;617:240–6.
Catalan G, Scott JF. Physics and applications of bismuth ferrite. Adv Mater. 2009;21:2463–85.
Chow CL, Ang WC, Tse MS, Tan OK. Oxygen-sensing property of sol-gel-derived
SrTi1xFexO3δ thin films with different iron concentrations (x = 0.2–0.8). Thin Solid Films.
2013;542:393–8.
Costa ACFM, Morelli MR, Kiminami RHGA. Combustion synthesis: effect of urea on the reaction
and characteristics of Ni-Zn ferrite powders. J Mater Synth Process. 2001;9:347–52.
Darshane S, Mulla IS. Influence of palladium on gas sensing performance of magnesium ferrite
nanoparticles. Mater Chem Phys. 2010;119:319–23.
Das S, Basu S, Mitra S, Chakravorty D, Mondal BN. Wet chemical route to transparent BiFeO3
films on SiO2 substrates. Thin Solid Films. 2010;518:4071–5.
Deraz NM. Fabrication, characterization and magnetic behaviour of alumina-doped zinc ferrite
nano-particles. J Anal Appl Pyrolysis. 2011;91:48–54.
Díaz-Castañón S, García-Zaldívar O, Faloh-Gandarilla J, Watts BE, Calderón-Piñar F, Hernández-
Landaverde MA, Espinoza-Beltran FJ. Synthesis of powders and thin films of bismuth ferrite
from solution: a magneto-electric study. Appl Phys A. 2014;117:1283.
Dumitrescu AM, Samoila PM, Nica V, Doroftei F, Iordan AR, Palamaru MN. Study of the
chelating/fuel agents influence on NiFe2O4 samples with potential catalytic properties. Powder
Technol. 2013;243:9–17.
Eason R. Pulsed laser deposition of thin films: applications-led growth of functional materials.
New York: Wiley; 2007.
Ederer C, Spaldin NA. Weak ferromagnetism and magnetoelectric coupling in bismuth ferrite. Phys
Rev B. 2005;71:060401(R).
Florea M, Alifanti M, Parvulescu VI, Mihaila-Tarabasanu D, Diamandescu L, Feder M, Negrila C,
Frunza L. Total oxidation of toluene on ferrite-type catalysts. Catal Today. 2009;141:361–6.
Gass J, Srikantha H, Kislov N, Srinivasan SS, Emirov Y. Magnetization and magnetocaloric effect
in ball-milled zinc ferrite powder. J Appl Phys. 2008;103:07B309–3.
Gharagozlou M. Influence of calcination temperature on structural and magnetic properties of
nanocomposites formed by Co-ferrite dispersed in sol-gel silica matrix using tetrakis
(2-hydroxyethyl) orthosilicate as precursor. Chem Cent J. 2011;5:1–7.
25 Ferrites Obtained by Sol-Gel Method 733

Ghasemi A. Development of uniaxial magnetocrystalline anisotropy in SrFe12x(DyGd)xO19 thin


film synthesized by incorporation of high coercivity nanoparticles in sol–gel method. J Magn
Magn Mater. 2014;361:112–7.
Gopalan EV, Al-Omari IA, Sakthi Kumar D, Yoshida Y, Joy PA, Anantharaman MR. Inverse
magnetocaloric effect in sol–gel derived nanosized cobalt ferrite. Appl Phys
A. 2010;99:497–503.
Gore SK, Mane RS, Naushad M, Jadhav SS, Zate MK, Alothmanc ZA, Hui BKN. Influence of Bi3+-
doping on the magnetic and Mössbauer properties of spinel cobalt ferrite. Dalton Trans.
2015;44:6384–90.
Gul IH, Pervaiz E. Comparative study of NiFe2xAlxO4 ferrite nanoparticles synthesized by
chemical co-precipitation and sol–gel combustion. Mater Res Bull. 2012;47:1353–61.
Haun JB, Yoon TJ, Lee H, Weissleder R. Magnetic nanoparticles biosensors. Wiley Interdiscip Rev
Nanomed Nanobiotechnol. 2010;2:291–304.
Hou X, Feng J, Liu X, Ren Y, Fan Z, Zhang M. Magnetic and high rate adsorption properties of
porous Mn1xZnxFe2O4 (0  x  0.8) adsorbents. J Colloid Interface Sci. 2011;353:524–9.
Hu P, Pan D, Wang X, Tian J, Wang J, Zhang S, Volinsky AA. Fuel additives and heat treatment
effects on nanocrystalline zinc ferrite phase composition. J Magn Magn Mater. 2011;323:569–73.
Ishikawa A, Tanahashi K, Futamoto M. Magnetic and structural properties of Ba–ferrite films
prepared by sol-gel processing. J Appl Phys. 1996;79:7080.
Jadhav SS, Shirsath SE, Patange SM, Jadhav KM. Effect of Zn substitution on magnetic properties
of nanocrystalline cobalt ferrite. J Appl Phys. 2010;108:093920.
Javed K, Li WJ, Ali SS, Shi DW, Khan U, Riaz S, Han XF. Enhanced exchange bias and improved
ferromagnetic properties in Permalloy-BiFe0.95Co0.05O3 core-shell nanostructures. Sci Rep.
2015;5:18203. https://doi.org/10.1038/srep18203.
Jing P, Du J, Wang J, Wei J, Pan L, Li J, Liu Q. Width-controlled M-type hexagonal strontium ferrite
(SrFe12O19) nanoribbons with high saturation magnetization and superior coercivity synthesized
by electrospinning. Sci Rep. 2015;5:15089. https://doi.org/10.1038/srep15089.
Kiran VS, Sumathi S. Comparison of catalytic activity of bismuth substituted cobalt ferrite
nanoparticles synthesized by combustion and co-precipitation method. J Magn Magn Mater.
2017;421:113–9.
Kumar A, Yadav KL. Magnetic, magnetocapacitance and dielectric properties of Cr doped bismuth
ferrite nanoceramics. Mater Sci Eng B. 2011;176:227–30.
Laokul P, Amornkitbamrung V, Seraphin S, Maensiri S. Characterization and magnetic properties of
nanocrystalline CuFe2O4, NiFe2O4, ZnFe2O4 powders prepared by the Aloe vera extract
solution. Curr Appl Phys. 2011;11:101–8.
Leu C, Lin T, Chen S, Hu C. Effects of bismuth oxide buffer layer on BiFeO3 thin film. J Am Ceram
Soc. 2015;98:724–31.
Li L. Glycol-assisted autocombustion synthesis of spinel ferrite CoFe2O4 nanoparticles: magnetic
and electrochemical performances. J Sol-Gel Sci Technol. 2011;58:677–81.
Li YQ, Huang Y, Qi SH, Niu FF, Niu L. Preparation, and magnetic and electromagnetic properties
of La-doped strontium ferrite films. J Magn Magn Mater. 2011;323:2224–32.
Li Y, Cooper JK, Liu W, Sutter-Fella CM, Amani M, Beeman JW, Javey A, Ager JW, Liu Y,
Toma FM, Sharp ID. Defective TiO2 with high photoconductive gain for efficient and stable
planar heterojunction perovskite solar cells. Nat Commun. 2016;7:12446. https://doi.org/
10.1038/ncomms12446.
Liang C, Yang D, Yang Z, Hou F, Xu M. The preparation and oxygen sensitivity of strontium ferrite
thin films. Surf Coat Technol. 2005;200(7):2515.
Lin Z, Cai W, Jiang W, Fu C, Li C, Song Y. Effects of annealing temperature on the microstructure,
optical, ferroelectric and photovoltaic properties of BiFeO3 thin films prepared by sol–gel
method. Ceram Int. 2013;39:8729–36.
Liu WT, Wu JM. The effect of the vacuum extraction and the Fe/Ba ratio on the phase formation of
barium ferrite thin film synthesized by sol–gel method. Mater Chem Phys. 2001;69:148–53.
Liu F, Yang C, Ren T, Wang AZ, Yu J, Liu L. NiCuZn ferrite thin films grown by a sol–gel method
and rapid thermal annealing. J Magn Magn Mater. 2007;309:75–9.
734 S. E. Shirsath et al.

Liu W, Tan G, Xue X, Dong G, Ren H, Xia A. Structure and multiferroic properties of Sr substituted
Bi0.89xSm0.11SrxFe0.94(Mn0.04Cr0.02)O3 thin films. Ceram Int. 2014;40:13249–56.
Liu W, Tan G, Xue X, Dong G, Ren H, Xia A. Conduction mechanisms and enhanced multiferroic
properties of 2-2 type Bi0.89Sm0.11FeO3–NiFe2O4 composite thin films. Ceram Int.
2015;41:1687–93.
Mathew DS, Juaug RS. An overview of structure and magnetism of spinel ferrite nanoparticles and
their synthesis in microemulsions. Chem Eng J. 2007;129:51–65.
Morisako A, Liu X, Matsumoto M, Naoe M. The effect of underlayer for Ba-ferrite sputtered films
on c-axis orientation. J Appl Phys. 1997;81:4374–6.
Mukherjee K, Majumder SB. Hydrogen sensing characteristics of nanocrystalline Mg0.5Zn0.5Fe2O4
thin film: effect of film thickness and operating temperature. Int J Hydrogen Energy.
2014;39:1185–91.
Nadeem K, Krenn H. Exchange bias, memory and freezing effects in NiFe2O4 nanoparticles.
J Supercond Nov Magn. 2011;24:717–20.
Nasir S, Asghar G, Malik MA, Anis-ur-Rehman M. Structural, dielectric and electrical properties of
zinc doped nickel nanoferrites prepared by simplified sol–gel method. J Sol-Gel Sci Technol.
2011;59:111.
Neel L. Magnetic properties of ferrites: ferrimagnetism and antiferromagnetism. Ann Phys.
1948;3:137–98.
Patange SM, Shirsath SE, Jangam GS, Lohar KS, Jadhav SS. Rietveld structure refinement, cation
distribution and magnetic properties of Al3+ substituted NiFe2O4 nanoparticles. J Appl Phys.
2011;109:053909.
Patil KC, Hegde MS, Rattan T, Aruna ST. Chemistry of nanocrystalline oxide materials: combus-
tion synthesis, properties and applications. Singapore: World Scientific Publishing; 2008.
Pechini MP. Patent US 3 330 697; 1967.
Pramanik NC, Fujii T, Nakanishi M, Takada J. Development of nanograined hexagonal barium
ferrite thin films by sol–gel technique. Mater Lett. 2005;59:468–72.
Quickel TE, Schelhas LT, Farrell RA, Petkov N, Le VH, Tolbert SH. Mesoporous bismuth ferrite
with amplified magnetoelectric coupling and electric field-induced ferrimagnetism. Nat
Commun. 2015;6:6562. https://doi.org/10.1038/ncomms7562.
Raja G, Gopinath S, Raj RA, Shukla AK, Alhoshan MS, Sivakumar K. Comparative investigation
of CuFe2O4 nano and microstructures for structural, morphological, optical and magnetic
properties. Physica E. 2016;83:69–73.
Rao CNR. Chemical approaches to the synthesis of inorganic materials. Wiley Easter/New Age
Internationals: New Delhi; 1993.
Rashad MM, Zaki ZI, El-Shall H. A novel approach for synthesis of nanocrystalline MgAl2O4
powders by coprecipitation method. J Mater Sci. 2009;44:2992–8.
Reitz C, Suchomski C, Haetge J, Leichtweiss T, Jaglicic Z, Djerdj I, Brezesinski T. Soft-templating
synthesis of mesoporous magnetic CuFe2O4 thin films with ordered 3D honeycomb structure
and partially inverted nanocrystalline spinel domains. Chem Commun. 2012;48:4471–3.
Salemizadeh S, Seyyed Ebrahimi SA. Influence of Fe/Ba ratio, basic agent and calcination
temperature on the physical properties of nanocrystalline barium hexaferrite thin films prepared
by a sol–gel method. J Non-Cryst Solids. 2009;355:982–5.
Sanchez C, Belleville P, Popall M, Nicole L. Applications of advanced hybrid organic–inorganic
nanomaterials from laboratory to market. Chem Soc Rev. 2011;40:696–753.
Seifikar S, Rawdanowicz T, Straka W, Quintero C, Bassiri-Gharb N, Schwartz J. Structural and
magnetic properties of sol–gel derived NiFe2O4 thin films on silicon substrates. J Magn Magn
Mater. 2014;361:255–61.
Sharifi I, Shokrollahi H, Doroodmand MM, Safi R. Magnetic and structural studies on CoFe2O4
nanoparticles synthesized by co-precipitation, normal micelles and reverse micelles methods.
J Magn Magn Mater. 2012;324:1854–186.
25 Ferrites Obtained by Sol-Gel Method 735

Shirsath SE, Mane ML, Yasukawa Y, Liu X, Morisako A. Chemical tuning of structure formation
and combustion process in CoDy0.1Fe1.9O4 nanoparticles: influence@pH. J Nanopart Res.
2013;15:1976.
Shirsath SE, Mane ML, Yasukawa Y, Liu X, Morisako A. Self-ignited high temperature synthesis
and enhanced super-exchange interactions of Ho3+-Mn2+-Fe3+-O2 ferromagnetic nano-
particles. Phys Chem Chem Phys. 2014;16:2347–57.
Shirsath SE, Liu X, Yasukawa Y, Li S, Morisako A. Switching of magnetic easy-axis using crystal
orientation for large perpendicular coercivity in CoFe2O4 thin film. Sci Rep. 2016;6:30074.
https://doi.org/10.1038/srep3007.
Slatineanu T, Diana E, Nica V, Oancea V, Caltun OF, Iordan AR, Palamaru MN. The influence of
the chelating/combustion agents on the structure and magnetic properties of zinc ferrite. Cent
Eur J Chem. 2012;10:1799.
Snelling EC. Soft ferrites/properties and application. 2nd ed. London: Butterworths; 1988.
Solovyova ED, Calzada ML, Belous AG. The effect of sol–gel preparation conditions on structural
characteristics and magnetic properties of M-type barium hexaferrite thin films. J Sol-Gel Sci
Technol. 2015;75:215–23.
Stoia M, Caizer C, Stefanescu M, Barvinschi P, Barbu-Tudoran L. Characterisation of nickel–zinc
ferrite/silica nanocomposites with low ferrite concentration obtained by an improved modified
sol–gel method. J Sol-Gel Sci Technol. 2011;58:126–34.
Sun HL, Shi H, Zhao F, Qi L, Gao S. Shape-dependent magnetic properties of low-dimensional
nanoscale Prussian blue (PB) analogue SmFe(CN)64H2O. Chem Commun. 2005;4339–4341.
https://doi.org/10.1039/b507240a.
Sun J, Wang Z, Wang Y, Zhu Y, Shen T, Pang L, Wei K, Li F. Synthesis of the nanocrystalline
CoFe2O4 ferrite thin films by a novel sol–gel method using glucose as an additional agent. Mater
Sci Eng B. 2012;177:269–73.
Szczygiel I, Winiarska K. Low-temperature synthesis and characterization of the Mn–Zn ferrite.
J Therm Anal Calorim. 2011;104:577–83.
Tang H, Zhang W, Peng B, Zhang W. Magnetically anisotropic barium ferrite thin films on sapphire
by sol–gel process. Thin Solid Films. 2010;518:3342–4.
Thakur S, Katyal SC, Gupta A, Reddy VR, Sharma SK, Knobel M, Singh M. Nickel-zinc ferrite
from reverse micelle process-structural and magnetic properties, Mössbauer spectroscopy
characterization. J Phys Chem C. 2009;113:20785–94.
Toksha BG, Shirsath SE, Mane ML, Patange SM, Jadhav SS, Jadhav KM. Autocombustion high-
temperature synthesis, structural and magnetic properties of CoCrxFe2xO4 (0  x  1.0).
J Phys Chem C. 2011;115:20905–12.
Vestal CR, Zhang ZJ. Synthesis and magnetic characterization of Mn and Co spinel ferrite-silica
nanoparticles with tunable magnetic core. Nano Lett. 2003;3:1739–43.
Waqas H, Qureshi AH. Low temperature sintering study of nanosized Mn–Zn ferrites synthesized
by sol–gel auto combustion process. J Therm Anal Calorim. 2010;100:529–35.
Yue Z, Guo W, Zhou J, Gui Z, Li L. Synthesis of nanocrystalline ferrites by sol gel combustion
process: the influence of pH variation. J Magn Magn Mater. 2004;270:216–23.
Zhang W, Li F, Tang H, Peng B, Zhang W. Morphology control of barium ferrite thin films.
Vacuum. 2012a;86:946–8.
Zhang W, Zhang W, Lu Z, Li F, Peng B. Growth of highly textured barium ferrite thin films on
sapphire and their magnetic properties. Thin Solid Films. 2012b;526:237–40.
Zhou B, Zhang YW, Liao CS, Cheng FX, Yan CH, Chen LY, Wang SY. Enhanced magneto-optical
Kerr effects and decreased Curie temperature in Co-Mn ferrite thin films. Appl Phys Lett.
2001;79:1849.
Zhou B, Zhang YW, Liao CS, Yan CH. Magnetism and phase transition for CoFe2xMnxO4
nanocrystalline thin films and powders. J Magn Magn Mater. 2002;247:70–6.

View publication stats

You might also like