Está en la página 1de 34

Author’s Accepted Manuscript

Characterization of SrCo1.5Ti1.5Fe9O19 hexagonal


ferrite synthesized by sol-gel combustion and solid
state route

R. Vinaykumar, R. Mazumder, J. Bera

www.elsevier.com/locate/jmmm

PII: S0304-8853(16)32216-8
DOI: http://dx.doi.org/10.1016/j.jmmm.2016.12.135
Reference: MAGMA62343
To appear in: Journal of Magnetism and Magnetic Materials
Received date: 15 September 2016
Revised date: 8 December 2016
Accepted date: 29 December 2016
Cite this article as: R. Vinaykumar, R. Mazumder and J. Bera, Characterization
of SrCo1.5Ti1.5Fe9O19 hexagonal ferrite synthesized by sol-gel combustion and
solid state route, Journal of Magnetism and Magnetic Materials,
http://dx.doi.org/10.1016/j.jmmm.2016.12.135
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Characterization of SrCo1.5Ti1.5Fe9O19 hexagonal ferrite synthesized by sol-gel combustion
and solid state route

R. Vinaykumar, R. Mazumder, J. Bera*

Department of Ceramic Engineering, National Institute of Technology, Rourkela 769008, India

_______________________________________________________

First author:

R. Vinaykumar,

E Mail: vinaykumar.r1984@gmail.com

Telephone: Mob: +91 7077109826

Department of Ceramic Engineering,

National Institute of Technology,

Rourkela, Odisha-769 008, India

*Corresponding author:

Japes Bera

Department of Ceramic Engineering,

National Institute of Technology,

Rourkela, Odisha-769 008, India.

E Mail: jbera@nitrkl.ac.in

Telephone: Mob: +91 9437246159

Highlights:
 SrCo1.5Ti1.5Fe9O19 ferrite was successfully prepared by sol–gel combustion process

using two types of raw materials for TiO2 namely titanium tetra-isopropoxide (TTIP)

and titanyl nitrate (TN).

 TN based sol-gel synthesis has been reported first time for the synthesis of Co-Ti

substituted SrM ferrite.

 Phase formation was comparatively easier in the TN-based sol-gel process.

 Compare to solid state route, the density, saturation magnetization, permeability and

permittivity were comparatively higher in the ferrite synthesized by chemical route

(TTIP and TN-based).

*Corresponding author. Tel.:+91- 0661-2462204, E-mail address: jbera@nitrkl.ac.in

Abstract

Co-Ti co-substituted SrM hexagonal ferrite (SrCo1.5Ti1.5Fe9O19) was synthesized by sol-gel

combustion and solid state route. The effects of sources of TiO2 raw materials; titanium tetra-

isopropoxide (TTIP) and titanyl nitrate (TN) on the phase formation behaviour and properties of

the ferrite were studied. The thermal decomposition behavior of the gel was studied using TG-

DSC. The phase formation behavior of the ferrite was studied by using X-ray powder diffraction

and FTIR analysis. Phase formation was comparatively easier in the TN-based sol-gel process.

The morphology of powder and sintered ferrite was investigated using scanning electron

microscope. Magnetic properties like magnetization, coercivity, permeability, tan δµ and

dielectric properties were investigated. The ferrite synthesized by sol-gel based chemical route
showed higher saturation magnetization, permeability and permittivity compared to the ferrite

synthesized by solid state route.

Keywords: Sol-gel synthesis; SrM Hexagonal ferrite; Co-Ti substitutions; Magnetic properties

1. Introduction

In the recent decade, researchers pointed out that there is an increase industry requirement of M-

type hexagonal ferrites because of its excellent magnetic properties, high electrical resistivity,

high Curie temperature, and better chemical stability [1]. These ferrites have become massively

important materials commercially and technologically for multitude of applications, such as

permanent magnets, magnetic recording and data storage materials, components for mobile and

wireless communications at microwave/GHz frequencies, electromagnetic (EM) wave absorbers

for EM interference, radar absorbing material (RAM) and stealth technologies [1-3].

SrFe12O19 (SrM), BaFe12O19 (BaM), and PbFe12O19 (PbM) ferrites are belonging to the family of

M-type hexagonal ferrites. They have a magnetoplumbite-type crystal structure with space group

P63/mmc, where Fe3+ ions are in three octahedral (12k, 4f2, and 2a), one tetrahedral (4f1) and one

five-fold coordination bipyramidal (2b) site, respectively [1, 4, 5]. The structure of SrM ferrite

can be described as an alternate stacking of four building blocks, namely S (spinel: [Fe3+6O8]2+),

R (hexagonal: [Sr2+Fe3+6O11]2-), S* and R*, where asterisk (*) indicates the rotation of

corresponding spinel and hexagonal blocks by 180° around the c axis.

Among M-type ferrites, the SrM ferrite has received a great deal of technological attention for

high-frequency application as it has higher values of magnetic properties than those of BaM or

PbM ferrites [1]. The pure SrM ferrite possesses large saturation magnetization (Ms: 74-92

emu/g), higher coercivity (Hc: 6.69 kOe) and high magnetocrystalline anisotropy constant (K1:

3.5 x106 erg/cm3) giving a large uniaxial crystalline anisotropy (Ha: 20 kOe) along the c-axis [4,
5]. Recently it has been reported that substituted hexagonal ferrites with planar anisotropy

exhibit excellent magnetic performance in the high-frequency region with higher ferromagnetic

resonance (FMR) frequencies compared to spinel ferrites. The FMR frequency of the ferrite can

also be varied by the substitution with different cations [6, 7]. The extensive investigation have

been carried out to modify the properties of SrM ferrites by substituting Fe3+ ions with trivalent

ions like Al3+, Ga3+, Cr3+, Mn3+ [8, 9] or with the pair substitution of divalent cations (Co2+, Zn2+,

Ni2+, Mg2+, etc) and tetravalent cations (Ti4+, Sn4+, Zr4+,Ir4+,etc) [10–13]. Among them,

substitution with Co2+Ti4+ pair is reported to alter the anisotropy and give rise to the noncollinear

magnetic structures [14-16]. Co2+cations prefer to occupy the 4f2 and 12k sites whereas,

nonmagnetic Ti4+cations prefer 12k site in M-type hexagonal structure [15, 16]. Ti4+ substitution

interrupts the magnetic interactions in a way that the two blocks SR and S*R* become more or

less magnetically decoupled. As a result, the magnetocrystalline anisotropy changes gradually

from uniaxial (along c axis) to planar (a-b plane) with increasing concentration of substitution.

Planar anisotropy in SrCoxTixFe12-2xO19 (SrCoTi-M) ferrite has been reported above a critical

Co-Ti concentration of x=1.1 to 1.3 and then the ferrite shows a typical soft magnetic behavior

[14,17,18]. Thompson et al. [17] reported that the FMR frequency of SrCoTi-M ferrite decreases

from ~50 GHz for x=0 (pure SrM) to below 1GHz for x=1.5 substitution concentration. This is

due to the weakening of crystal anisotropy field with increased substitution.

The conventional solid state route is the most extensively used method to produce SrCoTi-M

ferrite [19-22]. There are some deficiencies in this route such as the metal contamination from

the grinding media and necessary of high calcination temperatures more than 1200oC for a long

time, which lead to the formation of coarse and fused grains. However, the sol–gel combustion

synthesis has been demonstrated to be the optimal technique to synthesize ferrites with a

narrower grain-size distribution due to its atomic-level mixture of metal cations and low

crystallization temperature [23]. There are few reports available on sol–gel combustion synthesis
of BaM [24-28] and SrM [23, 29, 30] ferrite. However, there is only one report on the sol-gel

combustion synthesis of SrCoTi-M ferrite [17].

In this paper, we report the synthesis of SrCo1.5Ti1.5Fe9O19 ferrite by conventional solid state

method and sol–gel combustion method using two types of raw material for TiO2 namely

titanium tetra-isopropoxide and titanyl nitrate. The phase formation behavior, microstructure and

magnetic and dielectric properties of the ferrite were discussed.

2. Experimental procedure

2.1 Materials and Methods

Analytical grade Sr(NO3)2, Fe(NO3)3 . 9H2O, Co(NO3)3 . 6H2O, titanium tetra-isopropoxide

(C12H28O4Ti), titanyl nitrate (TiO(NO3)2), citric acid (C6H8O7) and ammonium hydroxide

(NH4OH) were used as starting raw materials for sol-gel combustion synthesis of

SrCo1.5Ti1.5Fe9O19 ferrite. Here two types of raw materials; titanium tetra-isopropoxide (TTIP)

and titanyl nitrate (TN) were used as TiO2 source in separate batches of experiments to study

the effect of the raw material on the phase formation and properties of the ferrite. The ferrite

synthesized using two respective TiO2 sources will be stated as TTIP and TN-based

powder/ferrite. The TN was prepared using the commercially available TiO2 through the process

described elsewhere [31]. The ferrite was also synthesized through conventional solid state

reaction route using SrCO3, Fe2O3, CoO, and TiO2 raw materials.

During sol-gel combustion synthesis of ferrite, the stoichiometric amount of all metal nitrates

were dissolved completely in distilled water to obtain solution-I. In the first experiment, required

molar ratio of TTIP was added dropwise to the solution-I and for the second experiment,

required amount of TN was added to the solution-I. The aqueous solution-II was prepared by

dissolving citric acid in distilled water. The solution I and II were then mixed with continuous
stirring for 30 min to form a transparent homogeneous solution. The molar ratio of citric acid to

total metal ions was maintained 1:1. Under this condition ammonium hydroxide was added

dropwise under the continuous stirring condition to adjust the pH of the solution to 7. The

produced sol was evaporated on a hot plate to transform into a gel. Further heating leads to a

self-propagating combustion of the gel with a violent exothermic reaction that propagated

through the entire gel in few seconds, and a foamy mass was swelled up. The foamy mass

yielded a loose agglomerated as burned powder when ground. The as burned powder was

calcined at a temperature ranging from 900 to 1200oC for 4h in an air atmosphere to study the

phase formation behavior of the ferrite.

In conventional solid state reaction route, the raw materials were ball milled for 20 h in a

polyethylene vial using zirconia balls and isopropyl alcohol as the milling media. The mixed

powder was dried and then calcined at 1200oC for 4h in air. The ferrite powders prepared by two

routes were pressed into pellet and toroid shape (outer diameter 13.0 mm, inner diameter 3.0

mm, and thickness 2.0 mm) using PVA binder. Finally, pressed specimens were sintered at

1300oC for 2h in air.

2.2 Characterization

The thermal decomposition behavior of TTIP and the TN-based gel was evaluated by

simultaneous analysis of thermogravimetry (TG) and differential scanning calorimetry (DSC)

using thermal analyzer (Model 449C; Netzsch, Germany) in the air atmosphere with the heating

rate of 10oC/min. The functional groups present in the dried gel, as burned and calcined powders

were identification by using Fourier Transformed Infrared (FTIR) spectroscopy (model 95277;

Perkin Elmer Spectrum) in the IR range 4000–400cm-1. The Phase formation behavior of the

ferrite was investigated by X-ray diffraction (XRD) analysis using the Rigaku diffractometer

(model; ULTIMA-IV/ Japan) with Cu Kα radiation. The Field Emission Scanning Electron
Microscope (FESEM Model; Nova NanoSEM/FEI) was used to analyze the morphology of

sintered ferrite. The magnetization behavior of the ferrite was measured using Vibrating Sample

Magnetometer (VSM; Lake Shore 7400) at a maximum applied field of 8 kOe.

The saturation magnetization (Ms) was deduced by a model described by R. Grossinger et al.

[32] as per the equation:

( ) (1)

Here, M is magnetization, H is magnetic field applied, b is a constant related with magneto-

crystalline anisotropy. Ms was calculated from the intercept of the straight line extrapolated in M

vs. 1/H2 plot for 1/H2 approaches zero and the slope was used to calculate b. The

magnetocrystalline anisotropy constant (K1) and the anisotropy field (Ha) were calculated using

equation (2) and (3) respectively [31]:

( ) (2)

(3)

The frequency dependent inductance, resistance, magnetic loss factor (tan δµ) were measured on

toroid samples with six turn wound by copper wire and capacitance, dielectric loss factor (tan δε)

were measured on pellet samples using 16192A parallel electrode test fixture attached with an

LCR meter (Model; Agilent E4982A) in the frequency range 1MHz to 100MHz at room

temperature. Before each measurement the LCR meter was calibrated using the 16195B

calibration kit for Open, Short, Low-Loss Capacitor, and 50 Ω Load. The complex initial

permeability was calculated from the inductance data. The real component of initial permeability

( was calculated as per the equation:


(4)

where L is the measured inductance of toroid, [ ] in which

N is the number of turns, Ht is the thickness, Do is the outer diameter and Din is the inner

diameter of the toroid sample. The imaginary component of initial permeability ( was

calculated as per the equation:

(5)

where R is the measured resistance of toroid, ω is angular frequency. The magnetic loss tangent

is also calculated as per the equation:

(6)

The real component of permittivity ( was calculated as per the equation:

(7)

where ε0 is the permittivity of free space, the C is capacitance, t and A are the thickness and

cross-sectional area of the sample. The imaginary component of permittivity ( was calculated

as per the equation:

(8)

3. Results and Discussions

3.1 Thermal decomposition behavior

Fig 1 (a) and (b) shows TG and DSC curves of TTIP and TN-based gel powders. For both the

gel, major weight loss was at 200oC. This weight loss corresponds to an enormous exothermic

peak in DSC curve. The exothermic peak at 200oC is due to the combustion reaction between

citric acid and nitrate ions in the gel [34]. The citric acid decomposes to CO2 and carbon
fragments, followed by the oxidation of carbon fragments with oxygen coming from NO-3 ion

through nitrate-citrate oxidation- reduction reaction.

In the case of TTIP based precursor the exothermic reaction almost complete at 200oC.

However, for TN based precursor there are two more exothermic peaks at 275 and 345oC

respectively. There was a weight loss under the sharp exothermic peak at 275oC, and there was

no weight loss under the broad exothermic peak at 345oC. The exothermic peak at 275oC

corresponds to the burning of residual carbon present in ash. The broad exothermic peak at

345oC may be due to the formation of crystalline phase from amorphous material [35]. The XRD

phase analysis showed the presence of two crystalline phases in TN-based as burned powder

(details of XRD phase analysis is presented in next section). There are some broad exothermic

peaks at 500oC and above in both TTIP and TN-based powder corresponding to the formation of

different crystalline phases. The crystalline phase formation is relatively slower process compare

to the oxidation- reduction reaction and the phase formation exotherms are broad in nature.

3.2 Phase formation behavior

Fig 2 shows the XRD patterns of TTIP based as burned and calcined powders. The as burned

powder XRD pattern is almost amorphous. The XRD patterns of powder after calcined at 900oC

for 4h shows the phases; strontium titanate (SrTiO3, JCPDS file number 79-0175) and γ-Fe2O3

(JCPDS file number 86-0550). In addition to these phases, CoFe2O4 (JCPDS file number 22-

1086) phase was found in the powder calcined at 1100oC for 4h. However, after calcination at

1200oC for 4h, the single phase SrM ferrite was found, and all the diffraction peaks were

indexed to M-type strontium hexagonal ferrite (JCPDS: 79-1412 ).

Fig 3 shows the XRD patterns of TN-based as burned and calcined powders. The as burned

powder XRD pattern shows the phases; SrFe0.5Ti0.5O2.85 (JCPDS file number 84-1004) and

CoFe2O4 (JCPDS file number 22-1086). The intermediate phase γ-Fe2O3 (JCPDS file number
86-0550) was found in the XRD patterns of powder calcined at 1000 oC for 4 h. However, after

calcination at 1100oC for 4h, a single SrM ferrite was found, and all the diffraction peaks were

indexed to M-type strontium hexagonal ferrite.

Fig.4 shows the XRD patterns of solid state, TTIP, and TN-based SrFe9Co1.5Ti1.5O19 sintered

ferrite. All the three samples exhibited a single phase M-type hexagonal ferrite, and there was no

secondary phase found in the XRD patterns. The Co–Ti substitution does not affect the basic

crystal structure of M-type hexaferrite because Ti4+ and Co2+ ions have an almost similar ionic

radius to that of Fe3+ ion.

3.3 FTIR analysis

In order to observe chemical and structural changes during synthesis of SrCo1.5Ti1.5Fe9O19

ferrite, FTIR spectra of powders were recorded. Fig 5 and 6 shows FTIR spectra of gel, as

burned powder and calcined powder synthesized using TTIP and TN-based raw materials

respectively. In the spectra of both TTIP and TN-based dried gel, the broad bands in the range

3139 - 3149 cm-1 are attributed to the characteristics O-H stretching vibration of the hydroxyl

group of citric acid. The bands in the range 1616-1624 cm-1and 1455 cm-1 indicate asymmetrical

and symmetrical stretching vibrations respectively of –COO- carboxyl group of citric acid. The

bands in the range 1762-1767cm-1and 2396-2398 cm-1 correspond to the free carboxyl group of

citric acid. The bands located at 1384cm-1and 825- 832 cm-1 range are assigned to N-O stretching

and bending vibration respectively of NO3- [25,34]. So the result indicates that the gel contains

the –COO- ,–COOH functional group of citric acid and nitrate anions.

In the spectra of as burned powder, the O-H stretching vibration band at 3401cm-1 is very

prominent in the case of TTIP based powder. However, the band is absent in TN-based powder.

The result indicates that the burning temperature in the case of the TN-based experiment was

higher than TTIP-based one. In case of TTIP based synthesis, combustion temperature was not
so high for complete removal of crystalline water. The bands at 586 and 456cm-1 are attributed

to spinel ferrite [35]. The XRD analysis of as burned TTIP and TN-based powders stated above

also showed the formation of CoFe2O4 spinel ferrite. The band at 1384cm-1 found in both TTIP

and TN-based as burned powder are assign to the stretching vibration of NO3- group remaining

in the as burned powders. The same observation of the presence of NO3- group in as burned

powder was reported in the literature [25, 34]. In addition, there is some vibration band at 1624

and 1455cm-1 corresponding to the carboxylic group are found in TTIP based as burned powder.

This again indicates that burning temperature in the TTIP based experiment was not sufficiently

high to remove all the organic components.

In the spectra of calcined powder, the bands at 586 cm-1 was identified as tetrahedral Fe-O

stretching and the bands in the range 432-430 cm-1were identified as tetrahedral Fe-O bending

and octahedral Fe-O stretching vibration of the M-type hexagonal ferrite [35-37]. These bands

are related to intrinsic stretching and bending vibration of the tetrahedral and octahedral Fe-O

bonds of hexagonal ferrite. Most of the –COO- , –COOH functional group and NO3- ion were

decomposed during calcination and no band related to these functional group are found in

calcined ferrite powder.

3.4 Surface morphology analysis

Fig 7. (a) and (b) shows FESEM images of TTIP and TN-based as burned ferrite powder. The

micrograph indicates that the grains are in nanometer scale. An average grain size of about 100

and 200 nm was found in TTIP and TN-based as-burned powder respectively. The TN-based

specimen shows the formation of well-developed crystalline phases. The bigger grain size and

their crystalline shape again proved that the combustion temperature was comparatively higher

in the TN-based experiment. Fig 8 (a), (b) and (c) shows FESEM images of solid state route,

TTIP, and TN-based sintered ferrite samples. The micrograph indicates that the plate-like grains
are with hexagonal shape and grains are in 3 to 5-micron size range. The thickness of plate-like

grains was higher in solid state specimen and lower in the TTIP-based specimen. The densities

of three sintered ferrite measured by Archimedes’s method is shown in Table 1. The density of

solid state method pellet (4.7 g/cm3) was lower than the density of TN and TTIP based samples

(4.95 g/cm3).

3.5 Magnetic properties analysis

The variation of magnetization as a function of applied field at room temperature is shown in

Fig.9. The magnetic properties of ferrite such as coercive field (Hc) and remanence

magnetization (Mr) were determined directly from the hysteresis loops. The saturation

magnetization was obtained from the intercept of the straight line of the M vs. 1/H2 plot (Inset

Fig. 9) in the field region 6 kOe < H < 8 kOe for all the samples as per equation (1). While Ha

and K1 were calculated according to the equations (2) and (3). Table 1 summarizes the magnetic

properties of three different route ferrite samples. The Ms was in the range of 44 to 54.1emu/g,

and Hc was in the range 20 to 55.3 Oe. Both Ms and Hc were comparatively lower than the pure

SrM ferrite. The magnetization decreased to 54.12 emu/g (TN-based) from 72 emu/g for pure

SrM ferrite. The ion replacement process plays a primary role in the decrease of the net

magnetization. The decrease in Ms was attributed to the substitution of higher magnetic moment

Fe3+ ions (5µB) by Ti4+ ion with zero magnetic moment and Co2+ ions with 3 µB moments. The

substitution also dilutes the strength of the Fe3+–O2-–Fe3+ exchange interactions and reduces the

total magnetic moment [38]. The coercivity decreased dramatically to 55.3 Oe (TN-based) from

6.69 kOe (pure SrM ferrite). It is known that the Co2+ and Ti4+ cations preferentially occupy 2b,

4f2 and 12k sites of M-type hexagonal crystal structure [15,16]. The origin of large MCA in pure

SrM ferrite has been explained by the presence of Fe ions primary on the 2b site of the

hexagonal structure [15]. Occupancy of the Co and Ti ions on the 2b site leads to the decrease in

magnetocrystalline anisotropy. For that reason, there is the overall reduction in coercivity as the
coercivity is closely related to the anisotropy fields in hexaferrite. The magnetization

characteristic showed typical soft magnet behavior of SrCo1.5Ti1.5Fe9O19 ferrite. This soft

property is attributed to the effect of Co-Ti pair substitution.

The magnetic moment (nB) in terms of μB was calculated by the relation; nB = (Molecular weight

of SrCo1.5Ti1.5Fe9O19 x Ms)/5585 as stated by the reference [9]. The values of nB and squareness

ratio (Mr/Ms) of the ferrite were about 9 μB and 0.1 respectively (shown in Table 1). As reported

in the literature, the pure SrM ferrite has nB of 20.6 μB and the squareness ratio (SR) of almost

0.5 [1]. The decreases in the values of nB to about 9 μB for the present ferrite samples is due to

the high-level of substitution of Fe3+ ions by non-magnetic Ti4+ and lower magnetic moment

Co2+ ions. Similarly, the decrease of SR to about 0.1 is due to the change of uniaxial anisotropy

to a planar anisotropy by the effect of Co-Ti substitution. The anisotropy constant (K1) and

anisotropy field (Ha) (shown in Table 1) were also found lower compared to pure SrM ferrite.

This was again due to the planar anisotropy in Co-Ti substituted SrM ferrite.

Fig. 10 shows the real (μ′) and imaginary (μ″) parts of the permeability respectively for different

ferrites in the frequency range from 1M to 100 MHz. The μ′, μ″ and tan δμ values at three

different frequencies are shown in Table 2. It can be seen from the Fig. 10 that the value of μ′

remains constant up to about 100 MHz frequency for all three samples. The permeability of solid

state, TTIP, and TN-based specimens were about 9, 15 and 16 respectively. The permeability is

closely correlated to the magnetization, densification and grain size of sintered ferrites [39]. The

TN and TTIP based sample shows higher μ′ compares to solid state sample due to its high Ms

and higher density. The imaginary (μ″) part of the permeability was about 0.4 for TN-based, 0.27

for TTIP-based and 0.1 for solid state based ferrite at 100 MHz frequency. The μ″ of TN-based

specimen was highest due to its highest μ′.

3.6 Dielectric properties analysis


Fig. 11 shows the real (ε′) and imaginary (ε ) parts of the permittivity respectively for

different ferrites in the frequency range from 1M to 100MHz. The ε′, ε and tan δε values at

three different frequencies are shown in Table 2. Both the ε′ and ε decreases with increasing

frequency. It is a typical behaviour of ferrites which can be explained with the help of Maxwell –

Wagener two layer model [40]. The decrease in permittivity with frequency might be originated

by the dielectric relaxation of the polarization which is generated by the space charges at the

grain boundary. The large value of permittivity at lower frequency is due to the predominance of

interfacial dislocation, oxygen vacancies, Fe2+ ions and grain boundary defect [41]. The ε′ was

comparatively higher in the ferrite synthesized by chemical route (TTIP and TN-based) compare

to solid state route. This may be due to the higher density and larger grain size of the ferrite

synthesized by chemical route. The ε values of all the samples are in the acceptable range.

The TN and TTIP based samples reveals the best performance compares to solid state sample

due to higher permeability and permittivity. The magneto-dielectric properties of the ferrite

suggest its high-frequency device applications.

4. Conclusions

SrCo1.5Ti1.5Fe9O19 was successfully prepared by sol–gel combustion method using two types of

raw material for TiO2 namely titanium tetra-isopropoxide (TTIP) and titanyl nitrate (TN). The

thermal decomposition behavior showed one stage and two stages weight losses respectively for

TTIP and TN-based gel. The pure SrCoTi-M ferrite phase was found after calcination of as

burned powder at 1100 and 1200oC respectively for TN and TTIP-based powder. This indicates

that the phase formation was easy in TN- based process. The FTIR functional group analysis

showed that the burning temperature in TTIP based combustion was not sufficiently high to

remove organic components completely. The microstructural analysis of as burned powder

showed the formation of plate-like crystalline phase in TN-based powder. All these indicate that

combustion temperature of TN-based process was higher than TTIP-based one. The density,
saturation magnetization, permeability, coercivity and permittivity were comparatively higher in

the ferrite synthesized by chemical route (TTIP and TN-based) compare to solid state route. The

ferrite synthesized by chemical route showed high permittivity of in the range 36 to 41, low tan

δε of ~0.014, high permeability in the range 15 to 17, low tan δμ of ~0.02 at the frequency 100

MHz. All these properties of the ferrite suggest for its application as an inductor material and in

antenna for very high frequency (VHF) band.

Acknowledgments

This work was supported by the SERB-DST (Grant No. SB/S3/ME/076/2013) of Government of

India. The authors would like to thank, Dr. Prasanta Chowdhury and Mr. Prabhanjan Kulkarni,

National Aerospace Laboratories, Bangalore, for the help on measuring the magnetic properties

of the samples.

References

[1] R.C. Pullar, Hexagonal ferrites: A review of the synthesis, properties and applications of
hexaferrites ceramics, Prog.Mater.Sci.57 (2012)191–1334. doi:10.1016/j.pmatsci.2012.04.001.

[2] M. Pardavi-Horvath, Microwave applications of soft ferrites, J. Magn. Magn. Mater. 215-216
(2000) 171-183.

[3] V.G. Harris, Modern microwave ferrites, IEEE Trans. Magn. 48 (2012) 1075–1104.
doi:10.1109/TMAG.2011.2180732.

[4] J. Smit, H.P.J. Wijn, Ferrites, Philips Technical Library, Eindhoven, 1959.

[5] P. Campbell, Permanent Magnet Materials, and Their Application, Cambridge University
Press, Cambridge, 1994.

[6] H.S. Cho, S.S. Kim, M-Hexaferrites with planar magnetic anisotropy and their application to
high-frequency microwave absorbers. IEEE Trans. Magn. 35 (1999) 3151.
[7] V.G. Harris, A. Geiler, Y. Chen, S.D. Yoon, M. Wu, A. Yang, et al., Recent advances in
processing and applications of microwave ferrites, J. Magn. Magn. Mater. 321 (2009) 2035–
2047. doi:10.1016/j.jmmm.2009.01.004.

[8] H. Luo, B.K. Rai, S.R. Mishra, V. V. Nguyen, J.P. Liu, Physical and magnetic properties of
highly aluminum doped strontium ferrite nanoparticles prepared by auto-combustion route, J.
Magn. Magn. Mater. 324 (2012) 2602–2608. doi:10.1016/j.jmmm.2012.02.106.

[9] B.K. Rai, S.R. Mishra, V.V. Nguyen, J.P. Liu, Synthesis and Characterization of high
Coercivity Rare-Earth ion Doped Sr0.9RE0.1Fe10Al2O19 (RE: Y, La, Ce, Pr, Nd, Sm, and Gd), J.
Alloys Compd. 550 (2012) 198–203. doi:10.1016/j.jallcom.2012.09.021.

[10] G. Mendozasuarez, Magnetic properties and microstructure of BaFe11.6-2xTixMxO19 (M=Co,


Zn, Sn) compounds, Phys. B Condens. Matter. 339 (2003) 110–118.
doi:10.1016/j.physb.2003.08.120.

[11] H. Sözeri, Z. Mehmedi, H. Kavas, A. Baykal, Magnetic and microwave properties of


BaFe12O19 substituted with magnetic, non-magnetic and dielectric ions, Ceram. Int. 41 (2015)
9602–9609. doi:10.1016/j.ceramint.2015.04.022.

[12] I. Bsoul, S.H. Mahmood, A.F. Lehlooh, A. Jamel, Structural and magnetic properties of
SrFe12-2x TixRuxO19, J. Alloys Compd. 551 (2013) 490–495.

[13] A. Ghasemi, A. Hossienpour, A. Morisako, X. Liu, A. Ashrafizadeh, Investigation of the


microwave absorptive behavior of doped barium ferrites, Mater. Des. 29 (2008) 112–117.
doi:10.1016/j.matdes.2006.11.019.

[14] J. Kreisel, H. Vincent, F. Tasset, M. Pate, J.P. Ganne, An investigation of the magnetic
anisotropy change in BaFe12-2x TixCoxO19 single crystals, J. Magn. Magn. Mater. 224 (2001) 17-
29.

[15] X. Batlle, X. Obradors, J. Rodríguez-Carvajal, M. Pernet, M. V. Cabañas, M. Vallet, Cation


distribution and intrinsic magnetic properties of Co-Ti-doped M-type barium ferrite, J. Appl.
Phys. 70 (1991) 1614–1623. doi:10.1063/1.349526.

[16] M.V. Cabanas, J.M. Gonzalez-Calbet, J. Rodriguez-Carjaval, M. Vallet-Regi , The solid


sloution BaFe12-xTixCoxO19 (x= 0-6) cationic distribution by neutron diffraction, J. Solid State
Chem. (1994)111-229.
[17] S. Thompson, N.J. Shirtcliffe, E.S. O’Keefe, S. Appleton, C.C. Perry, Synthesis of
SrCoxTixFe(12−2x)O19 through sol–gel auto-ignition and its characterisation, J. Magn. Magn.
Mater. 292 (2005) 100–107. doi:10.1016/j.jmmm.2004.10.102.

[18] D.J. De Bitetto, Anisotropy fields in hexagonal ferrimagnetic oxides by ferrimagnetic


resonance, J. Appl. Phys. 35 (1964) 3482–3487. doi:10.1063/1.1713254.

[19] L. Wang, D. Wang, Q. Cao, Y. Zheng, H. Xuan, J. Gao, et al., Electric control of
magnetism at room temperature, Sci. Rep. 2 (2012) 1–5. doi:10.1038/srep00223.

[20] E. P. Naiden,V. A. Zhuravlev, V. I. Itin, R. V. Minin, V. I. Suslyaev,O. A. Dotsenko,


Structure and Static and Dynamic Magnetic Properties of Sr(CoxTix)Fe12–2xO19 Hexaferrites
produced by Self-Propagating High-Temperature Synthesis, Russian Physics Journal, Vol. 55,
No. 8, ( 2013) 869-877.

[21] M.R. Eraky, Electrical conductivity of cobalt titanium substituted SrCaM hexaferrites, J.
Magn. Magn. Mater. 324 (2012) 1034–1039. doi:10.1016/j.jmmm.2011.10.021.

[22] C. Singh, S.B. Narang, I.S. Hudiara, Y. Bai, K. Marina, Hysteresis analysis of Co-Ti
substituted M-type Ba-Sr hexagonal ferrite, Mater. Lett. 63 (2009) 1921–1924.
doi:10.1016/j.matlet.2009.06.002.

[23] M.G. Hasab, S.A.S. Ebrahimi, A. Badiei, Effect of different fuels on the strontium
hexaferrite nanopowder synthesized by a surfactant-assisted sol-gel auto-combustion method, J.
Non. Cryst. Solids. 353 (2007) 814–816. doi:10.1016/j.jnoncrysol.2006.12.048.

[24] S. Chaudhury, S.K. Rakshit, S.C. Parida, Z. Singh, K.D.S. Mudher, V. Venugopal, Studies
on structural and thermo-chemical behavior of MFe12O19 (M = Sr, Ba and Pb) prepared by
citrate-nitrate gel combustion method, J. Alloys Compd. 455 (2008) 25–30.
doi:10.1016/j.jallcom.2007.01.075.

[25] J. Huang, H. Zhuang, W. Li, Synthesis and characterization of nano crystalline BaFe12O19
powders by low temperature combustion, Mater. Res. Bull. 38 (2003) 0–10. doi:10.1016/S0025-
5408(02)00979-0.

[26] A. Mali, A. Ataie, Influence of the metal nitrates to citric acid molar ratio on the
combustion process and phase constitution of barium hexaferrite particles prepared by sol-gel
combustion method, Ceram. Int. 30 (2004) 1979–1983. doi:10.1016/j.ceramint.2003.12.178.
[27] G. Xu, H. Ma, M. Zhong, J. Zhou, Y. Yue, Z. He, Influence of pH on characteristics of
BaFe12O19 powder prepared by sol-gel auto-combustion, J. Magn. Magn. Mater. 301 (2006) 383–
388. doi:10.1016/j.jmmm.2005.07.014.

[28] A. Mali, A. Ataie, Influence of Fe/Ba molar ratio on the characteristics of Ba-hexaferrite
particles prepared by sol-gel combustion method, J. Alloys Compd. 399 (2005) 245–250.
doi:10.1016/j.jallcom.2005.03.017.

[29] M. Ghobeiti Hasab, S.A. Seyyed Ebrahimi, A. Badiei, An investigation on physical


properties of strontium hexaferrite nanopowder synthesized by a sol-gel auto-combustion
process with addition of cationic surfactant, J. Eur. Ceram. Soc. 27 (2007) 3637–3640.
doi:10.1016/j.jeurceramsoc.2007.02.004.

[30] C.S. Lin, C.C. Hwang, T.H. Huang, G.P. Wang, C.H. Peng, Fine powders of SrFe12O19
with SrTiO3 additive prepared via a quasi-dry combustion synthesis route, Mater. Sci. Eng. B
Solid-State Mater. Adv. Technol. 139 (2007) 24–36. doi:10.1016/j.mseb.2007.01.053.

[31] R. Mazumder, A. Sen, “Ultra”-low-temperature sintering of PZT: A synergy of nano-


powder synthesis and addition of a sintering aid, J. Eur. Ceram. Soc. 28 (2008) 2731–2737.
doi:10.1016/j.jeurceramsoc.2007.12.032.

[32] R. Grossinger, critical Examination of the Law of Approach to Saturation, Phys. stat. sol. a

66 (1981) 665-674.

[33] S. Güner , I.A. Auwal , A. Baykal , H. Sözeri, Synthesis, characterization and magneto

optical properties of BaBixLaxYxFe12-3xO19 hexaferrites, J. Magn. Magn. Mater. 416 (2016) 261–

268. doi.org/10.1016/j.jmmm.2016.04.091

[34] Y.S. Hong, C.M. Ho, H.Y. Hsu, C.T. Liu, Synthesis of nanocrystalline Ba(MnTi)xFe12-2xO19
powders by the sol-gel combustion method in citrate acid-metal nitrates system (x=0, 0.5, 1.0,
1.5, 2.0), J. Magn. Magn. Mater. 279 (2004) 401–410. doi:10.1016/j.jmmm.2004.02.008.

[35] S.M. Masoudpanah, S.A.S. Ebrahimi, Structure and magnetic properties of nanocrystalline

SrFe12O19 thin films synthesized by the Pechini method, J. Magn. Magn. Mater. 342 (2013) 128–

133. doi:10.1016/j.jmmm.2013.04.070.
[36] W. Zhao et al., FTIR Spectra, Lattice Shrinkage, and Magnetic Properties of CoTi-

Substituted M-Type Barium Hexaferrite Nanoparticles, J. Am. Ceram. Soc., 90 (2007) 2095–

2103.

[37] S.K. Chawla, S.S. Meena, P. Kaur, R.K. Mudsainiyan, S.M. Yusuf, Journal of Magnetism

and Magnetic Materials Effect of site preferences on structural and magnetic switching

properties of Co–Zr doped strontium hexaferrite SrCoxZrxFe(12-2x)O19, J. Magn. Magn. Mater.

378 (2015) 84–91. doi:10.1016/j.jmmm.2014.10.168.

[38] W. Zhang, Y. Bai, X. Han, L. Wang, X. Lu, L. Qiao, Magnetic properties of Co – Ti

substituted barium hexaferrite, J. Alloys Compd. 546 (2013) 234–238.

doi:10.1016/j.jallcom.2012.08.029.

[39] A. Globus, Magnetization Mechanisms: some Physical Considerations about the domain

wall Size theory of Magnetization Mechanisms, Le J. Phys. Colloq. 38 (1977) C1-1-C1-15.

doi:10.1051/jphyscol:1977101.

[40] C.G. Koops, On the Dispersion of Resistivity and Dielectric Constant of Some

Semiconductors at Audio frequencies, Phys. Rev. 83 (1951) 121.

[41] Z. Haijun, L. Zhichao, M. Chenliang, Y. Xi, Z. Liangying, W. Mingzhong, Preparation and

microwave properties of Co- and Ti-doped barium ferrite by citrate sol-gel process, Mater.

Chem. Phys. 80 (2003) 129–134. doi:10.1016/S0254-0584(02)00457-1.


Fig.1 TG-DSC curve of (a) TTIP and (b) TN-based SrCo1.5Ti1.5 Fe9O19 hexagonal ferrite.

Fig.2 Room temperature XRD pattern of TTIP based ferrite powder (a) As burned and after

calcinations at (b) 900 (c) 1100 and (d) 1200oC.

Fig.3 Room temperature XRD pattern of TN based ferrite powder (a) As burned and after

calcinations at (b) 900 (c) 1000 and (d) 1100oC.

Fig.4 XRD pattern of (a) solid state (b) TTIP and (c) TN-based sintered ferrites obtained at

1300oC.

Fig.5 FTIR spectra of TTIP based (a) dried gel (b) as burned and (c) 1200oC calcined ferrite

powders.

Fig.6 FTIR spectra of TN-based (a) dried gel (b) as burned and (c) 1100oC calcined ferrite

powders.

Fig.7 FESEM images of (a) TTIP and (b) TN-based as burned powder.

Fig.8 FESEM images of (a) TN (b) TTIP and (c) solid state route synthesized sintered ferrite.

Fig.9 Magnetic hysteresis loops of ferrite samples synthesized by (a) solid state route and by sol-

gel route using (b) TTIP and (c) TN based raw materials. The inset figure shows the plot of Ms

as a function of 1/H2 for TN based ferrite.


Fig.10 Frequency dependence complex permeability (μ and μ″) spectra of SrCo1.5Ti1.5Fe9O19

ferrites synthesized by (a) solid state route and by sol-gel route using (b) TTIP and (c) TN based

raw materials.

Fig.11 Frequency dependence complex permittivity (ε′ and ε ) spectra of SrCo1.5Ti1.5Fe9O19

ferrites synthesized by (a) solid state route and by sol-gel route using (b) TTIP and (c) TN based

raw materials.

Table 1 Density (D), saturation magnetization (Ms), remanence magnetization (Mr), coercivity

(Hc), squareness ratio (Mr/Ms), magnetic moment (nB), the anisotropy constant (K1) and

anisotropy field (Ha) of SrCo1.5Ti1.5Fe9O19 ferrite samples.

D Ms Mr Hc nB Mr/Ms K1 Ha
(gm/cm3) (emu/gm) (emu/gm) (Oe) (µB) (x104 (kOe)
Sample
erg/cm3)

TN 4.95 54.1 6.0 55.3 10.2 0.11 9.3 4.21


based

TTIP 4.95 52.6 5.2 46.2 10 0.98 9.3 3.54


based
solid 4.7 44 1.9 20 8.3 0.05 8.9 3.28
state
Table 2 Complex permeability (μ′, μ″), magnetic loss tangent (tan δμ), complex permittivity (ε′,

ε″) and dielectric loss tangent (tan δε) at 1, 10, and 100 MHz for three SrCo1.5Ti1.5Fe9O19 ferrite

samples.

Frequency Complex permeability tan δμ Complex permittivity tan δε


Sample
(MHz) μ′ μ″ ε′ ε″
TN 1 17.1 0.04 0.002 169 5.4 0.30
based 10 16.5 0.35 0.021 126 0.89 0.03
100 17 0.4 0.02 41 0.57 0.014
TTIP 1 15.6 0.11 0.007 160 9.9 1.41
based 10 14.8 0.26 0.017 124 5.6 0.14
100 15.3 0.27 0.017 37 0.55 0.023
solid 1 9 0.17 0.02 72 2.2 0.03
state 10 8.3 0.20 0.024 37 0.73 0.02
100 8.5 0.10 0.013 12 0.24 0.02

También podría gustarte