2015_Effects of Heat Treatment on the Phase Evolution, Structural

Solid State Phenomena Vol 232 (2015) pp 65-92
© (2015) Trans Tech Publications, Switzerland
doi:10.4028/www.scientific.net/SSP.232.65
Effects of Heat Treatment on the Phase Evolution, Structural, and
Magnetic Properties of Mo-Zn Doped M-type Hexaferrites
Sami H. Mahmood1,a*, Aynour N. Aloqaily1,b, Yazan Maswadeh1,c,
Ahmad Awadallah1,d, Ibrahim Bsoul2,e, Mufeed Awawdeh3,f, Hassan Juwhari1,g
1
Physics Department, University of Jordan, Amman 11942, Jordan
Physics Department, Al al-Bayt University, Mafraq 13040, Jordan
3
Physics Department, Yarmouk University, Irbid 21163, Jordan
2
a*
[email protected] (corresponding author), [email protected],
[email protected], [email protected], [email protected],
f
[email protected], [email protected]
c
Key Words: M-type Hexaferrite, W-type Hexaferrite, Structural Properties, Magnetic Properties,
Mössbauer Spectroscopy
Abstract. In this article we report on the structural and magnetic properties of BaFe12-4xMoxZn3xO19
hexaferrites with Mo-Zn substitution for Fe ions. The starting materials were commensurate with
the BaM stoichiometry, and the Mo:Zn ratio was 1:3. The powder precursors were prepared by high
energy ball milling, and subsequently sintered at temperatures from 1100 to 1300° C. The structural
analyses indicated that all samples sintered at 1100° C were dominated by a major M-type
hexaferrite phase. The relative abundance of the BaMoO4 and Zn-spinel secondary phases increased
with increasing the concentration of the substituents, resulting in a decrease of the saturation
magnetization from about 67 emu/g (for x = 0.0) to 55 emu/g (for x = 0.3). The coercivity also
decreased from 3275 Oe (for x = 0.0) to 900 Oe (for x = 0.3), demonstrating the ability to tune the
coercivity to the range useful for magnetic recording by the substitution process. The saturation
magnetization improved significantly with sintering at T > 1100° C, and the coercivity decreased
significantly, signaling the transformation of the samples to soft magnetic materials. These
magnetic changes were due to the high-temperature reaction of the spinel phase with the BaM phase
to produce the W-type hexaferrite phase on the one hand, and to the growth of the particles on the
other hand. The magnetic phases were further investigated using Mössbauer spectroscopy and
thermomagnetic measurements. Our study indicated that the sample with x = 0.2 has the highest
saturation magnetization (74 emu/g at sintering temperature of 1300° C) and a tunable coercivity
between 2100 Oe and 450 Oe.
Contents of Paper
1. Introduction
2. Properties of Hexaferrites
2.1. M-Type Hexaferrites
2.2. W-Type Hexaferrites
2.3. The MoZn-substituted M-Type Hexaferrites
3. Experimental Techniques
3.1. Sample Preparation
3.2. Sample Characterization
4. Results and Discussion
4.1. Scanning Electron Microscopy
4.2. X-Ray Diffraction
4.3. Mössbauer Spectroscopy
4.4. Magnetic Measurements
4.4.1. Hysteresis Characteristics
4.4.2. Initial Magnetization Curves
4.4.3. Thermomagnetic Measurements
All rights reserved. No part of contents of this paper may be reproduced or transmitted in any form or by any means without the written permission of Trans
Tech Publications, www.ttp.net. (ID: 86.108.122.87-20/04/15,09:25:02)
66
Ferroic Materials: Synthesis and Applications
5. Conclusions
References
1. Introduction
The discovery of barium hexaferrites (M-type or BaM hexaferrite) more than six decades ago, and
their promising wide range of technological and industrial applications, had generated great interest
in the fabrication and characterization of this class of ferrites [1 – 5]. The progress in manufacturing
microwave (MW) devices for different applications such as in home appliances, radar and
telecommunication systems had driven a great attention to the problem of electromagnetic
interference and radiation pollution. Consequently, interest was developed in the production of
magnetic materials for MW applications, and different members of the hexagonal ferrites, namely,
the W-type, the Y-type, the X-type, the Z-type, and the U-type, were successfully synthesized and
characterized by different techniques [1, 3, 6 – 9]. The wide range of applications of hexaferrites is
facilitated by the fact that their magnetic properties cover the range from those characteristic of soft
magnetic materials to those of hard magnetic materials. Magnetically hard materials are used for the
production of permanent magnets for motors, home appliances, and vehicles, while soft magnetic
materials are used in transformers, power supplies, and electronic components for
telecommunication applications, and hexaferrites with soft and intermediate magnetic hardness are
used in magnetic recording as components for read-write heads and data storage [10 – 13].
The major contribution to the coercivity of the hexaferrites is due to magnetocrystalline anisotropy
arising from the coupling of the spins of Fe3+ at different sites. According to Stoner-Wohlfarth
model for a random assembly of uniaxial single-domain barium hexaferrites particles, the coercivity
is given by [14]:
𝐻𝑐 =
0.96 𝐾1
𝑀𝑠
(1)
Here K1 is the first anisotropy constant and Ms is the saturation magnetization. In addition, shape
anisotropy plays a secondary role in determining the coercivity of these hexaferrites. The high
coercivity of barium hexaferrites (typically in the range of 2000 – 4000 Oe) is prohibitive for use in
data storage media as conventional heads are incapable of swithing the particles. Reduced
coercivities for high density magnetic recording media or other applications requiring lower
coercivities can be achieved by partial substitution of Fe3+ ions by specific combinations of Mn2+,
Ni2+, Mg2+, Co2+, Zn2+, Ti2+ or Sn2+ ions with tetravalent ions such as Ti4+, Ru4+, or Sn4+ [15-27].
On the other hand, improvement of the coercivity for applications requiring hard magnetic
materials was achieved by substituting Fe3+ ions by a trivalent metal ion such as Mn3+, Al3+, Ga3+,
In3+, Sc3+, As3+, and Cr3+ [28 – 31]. In addition, the structural and physical properties of hexaferrite
powders were modified by by adopting different physical or chemical preparation methods, as well
as varying the experimental conditions such as stoichiometry, heat treatment, and binder additives
[31-39].
Mössbauer spectroscopy was used to obtain information on the local chemical and structural
environment in hexaferrites. This information was employed to determine local structural
symmetries, preferential site ocupation, and valence state of Fe ions in the hexaferrite lattice [5, 15
– 17, 23, 28, 40, 41]. However, clear differences in the reported hyperfine parameters of
hexaferrites were found in the literature [41 – 47]. These differences could be due to different
nature of samples prepared by different methods and under different experimental conditions.
Further, the complexity of Mössbauer spectra and similarity of the hyperfine parameters
corresponding to some crystallographic sites could lead to some confusion in interpreting the
Mössbauer spectra.
Solid State Phenomena Vol. 232
67
2. Properties of Hexaferrites
2.1. M-Type Hexaferrites: M-type barium hexaferrite (BaFe12O19) is a hard ferrimagnetic material
with Curie temperature 450° C, melting point of 1390 °C, molecular mass of about 1112 g, and
maximum density of ρ = 5.28 g/cm3. It has a hexagonal structure with typical lattice parameters a =
b = 5.88 Å and c = 23.2 Å [1]. The unit cell of M-type barium hexaferrite consists of two R blocks
and two S blocks in the stacking sequence RSR*S*, where the star denotes a block rotated by 180°
about the c-axis of the hexagonal lattice. The S (Fe6O8) block contains two hexagonal layers of
oxygen, while the R block (BaFe6O11) consists of three hexagonal oxygen layers, with one Ba
cation replacing an oxygen anion substitutionally in the middle layer. The unit cell therefore
contains two (BaFe12O19) molecules. In the BaM lattice, there are five different interstitial sites
where small metal ions (such as Fe ions) reside. One octahedrally coordinated (2a) site and two
tetrahedrally coordinated (4f1) sites exist in each S block, while two octahedrally coordinated (4f2)
sites and one five-fold coordinated (2b) bi-pyramidal site exist in each R block. In addition, there
are three octahedral (12k) sites in each interface between an S and an R block. BaM is a uniaxial
crystal with a c-easy axis. The spins of the magnetic ions at 2a, 2b, and 12k sites are parallel to the
c-axis, forming three spin-up sublattices, while the spins of the ions at the 4f1 and 4f2 sites are
aligned antiparallel to the c-axis, thus forming two spin-down sublattices in the hexaferrite
structure. The characteristics of these sublattices, and their positions in the hexaferrite lattice are
shown in Table 1.
Table 1. Metallic sub-lattices of M-type hexaferrite
Block
sublattice
coordination
ions per unit Spin
cell
direction
S
4f1
Tetrahedral
4
down
2a
Octahedral
2
Up
4f2
Octahedral
4
Down
2b
Bi-pyramidal
2
Up
12k
Octahedral
12
Up
R
S-R
Assuming spin collinear structure in barium hexaferrites (at 0 K), the net magnetic moment per unit
cell of BaM is 20 µB, which corresponds to a saturation magnetization of about 100 emu/g [48].
This theoretical value is consistent with low-temperature magnetization measurements [1, 26]. The
saturation magnetization of BaM at 20 K is 72 emu/g, and its Curie temperature is 450° C.
2.2. W-Type Hexaferrites: The W-type hexaferrite with chemical formula BaMe2Fe16O27 (Me is a
divalent metal ion such as Co, Zn, Ni, Mg, and Mn) is a ferrimagnetic material with Curie
temperature ranging from 415° C (for Mn2W) to 455° C (for Fe2W) [48]. ZnFe-W has a Curie
temperature of 430° C, a theoretical density of about 5.31 g/cm3, and a saturation magnetization of
108 emu/g at 0 K and 73 emu/g at 20° C [1, 48]. The lattice parameters of the W-type are typically
a = b = 5.88 Å and c = 32.8 Å [1]. The unit cell of the W-type hexaferrite is built from double
blocks S and blocks R in accordance with the stacking sequence RSSR*S*S*. In this phase there
are seven distinct crystallographic sites: two tetrahedral (4e and 4fIV), four octahedral (4fVI, 6g, 4f,
and 12k) and one bi-pyramidal (2d) sites (see Table 2).
68
Ferroic Materials: Synthesis and Applications
Table 2. Iron sites, their coordination, spin orientation, number of Fe ions per formula, and
location in the W-type lattice.
Sub-lattice
Coordination
Location
Number of Spin
ions
orientation
4e
tetrahedral
S
2
down
4fIV
tetrahedral
S
2
down
4fVI
octahedral
R
2
down
4f
octahedral
S
2
up
2d
Bi-pyramidal
R
1
up
12k
octahedral
R-S
6
up
6g
octahedral
S-S
3
up
2.3.
The MoZn-Substituted M-Type Hexaferrite: While BaM hexaferrites with various
substitutions for Fe ions occupied a large area of the studies involving this important class of
magnetic materials, substitutions involving molybdenum were found to be extremely rare in the
literature. In a previous study, Mössbauer spectroscopy was reported on CoMo substituted M-type
hexaferrites, where it was found that the substitution occurs at the 12k sites [49]. This study,
however, was limited to Mössbauer spectroscopy and did not address the details of the structural
and magnetic properties of the ferrites.
In a series of recent studies performed in our laboratories, the effects of MoZn substitutions under
different experimental conditions were investigated. [50 – 52]. The magnetic properties and
hyperfine parameters were reported for BaFe11.6MoxZn0.4-xO19 (x = 0.1, 0.2, 0.4) hexaferrites with
Ba/Fe ratio of 1:11 prepared by wet chemical mixture method and sintered at 1100° C, with the
focus on the distribution of Mo ions having different valence states (Mo4+ and Mo6+) and its effects
on the magnetic properties [50]. In this study, Mo4+ and Mo6+ ions were reported to occupy (4f1 +
12k) and 2b sites, respectively. The sample with x = 0.4 demonstrated preference of the Mo4+ and
Mo6+ ions for 2b and 12k sites, with the development of the Fe2+ valence state. These ferrites
exhibited an increase in saturation magnetization which culminated at x = 0.2, and then a decrease
at x = 0.4. The saturation magnetization for these compounds were relatively high (70 – 75 emu/g),
and the coercivity changed from about 1980 Oe to 2600 Oe, which are properties of materials with
potential for high density magnetic recording [50]. In addition, the magnetic properties and
hyperfine interactions of BaFe12-2xMoxZnxO19 (0.0 < x < 0.3) hexaferrites with Ba/Fe ratio of 1:11
prepared by wet chemical mixture method and sintered at 1100° C, were investigated [51]. The
coercivity of this system was found to change from 2600 Oe (for x = 0.0) to about 1060 Oe (for x =
0.3) while the saturation magnetization increased from 71 emu/g for the un-doped sample to 75
emu/g for the sample with x = 0.15. This system also demonstrated tunable magnetic properties for
potential high density magnetic recording.
On the other hand, BaFe12-4xMoxZn3xO19 (0.0 < x < 0.3) hexaferrites with Ba/Fe ratio of 1:11
prepared by ball milling and sintered at temperatures between 1100° C and 1300° C were studied
[52]. The saturation magnetization for the samples sintered at 1100° C decreased from about 70
emu/g for the un-doped sample down to 57 emu/g for the sample with x = 0.3, and the decrease was
attributed to the presence of nonmagnetic impurity phases. These samples demonstrated hard
magnetic properties with coercivity variations between about 2600 Oe (for x = 0.1) and 3450 Oe
(for x = 0.0). Upon increasing the sintering temperature, however, the saturation magnetization of
the sample with x = 0.2 improved significantly (58.9 emu/g for the sample sintered at 1100° C, and
Solid State Phenomena Vol. 232
69
70.5 emu/g for the sample sintered at 1300° C), and the coercivity decreased significantly (from
about 2800 Oe to about 300 Oe). These dramatic changes were associated with the development of
the W-type hexaferrite phase at high sintering temperatures.
In view of the above discussion, the MoZn-substituted barium hexaferrite system seems to have rich
structural and magnetic phase diagram, and provides the opportunity to produce materials with
tunable magnetic properties. Therefore, in this article, we report on the structural and magnetic
properties of BaFe12-4xMoxZn3xO19 hexaferrite prepared by ball milling and different experimental
conditions. The Mo:Zn ratio of 1:3 was chosen to ensure that molybdenum ions in the hexaferrite
lattice would have the Mo6+ valence state. The structural properties and the particle morphology of
the prepared compounds were investigated by powder x-ray diffraction (XRD) and scanning
electron microscopy (SEM), respectively. Also, the effects of Mo6+Zn2+ concentration on the phase
evolution, cationic distribution, and the magnetic properties were investigated by Mössbauer
spectroscopy, and isothermal and thermomagnetic measurements.
3. Experimental Techniques
3.1. Sample Preparation: Powder precursors of barium hexaferrites with stoichiometry consistent
with the chemical formula BaFe12-4xMoxZn3xO19 were prepared from high purity (~99%) powders of
BaCO3, Fe2O3, ZnO and MoO2 (Sigma-Aldrich made). The barium to metal molar ratio was 1:11.5,
and the mixing and homogenization of the powder precursors was achieved by using a planetary
ball mill (Fritch Pulverisette 7). The milling cups and balls were made of zirconia, and the powderto-ball mass ratio was 1:14 to ensure effective grinding and obtain homogenous powders. Wet
milling of the powder precursors in an acetone bath (8 ml for each 5 grams of the powder) was
performed for 16 hours at a rotational speed of 250 rpm. The wet powder was then left in the
container overnight to dry at room temperature, and subsequently collected in clean glass vials.
About 1 g of the powder was mixed with an adhesive agent of 2% wt. of polyvinyl alcohol (PVA)
and then dry-pressed into disc-shape pellet (1.5 cm in diameter) in a stainless steel die under a 4 ton
force. The discs were then sintered at temperatures ranging from 1100° C to 1300° C for 2 h in air.
3.2. Sample Characterization: The prepared samples were characterized using standard structural
and magnetic measurements. X-ray diffraction (XRD) was used to determine the structural phases
of the sample, and examine their characteristics and abundance. XRD patterns were collected in the
angular range 20° ≤ 2θ ≤ 70° in steps of 0.02° using an XRD 7000-Shimadzu diffractometer with
Cu-Kα radiation. The patterns were analyzed by the Expert High Score software to identify the
structural phases, and then Rietveld refinement was performed to fit the patterns using the FullProf
software [52]. The particle morphology and particle size distribution were examined by scanning
electron microscopy using FEI-Inspect F50/FEG instrument.
Mössbauer spectroscopy was performed using a conventional constant acceleration spectrometer
with 57Co/Cr source. The spectrum of metallic iron at room temperature was used to calibrate the
energy scale, and the isomer shift was determined with respect to its center. The spectra were
collected over 512 channels, and then analyzed using a standard fitting routine based on least square
minimization technique.
The magnetic measurements were conducted using a vibrating sample magnetometer
(VSMMicroMag 3900, Princeton Measurements Corporation). Magnetization measurements were
made at different temperatures in applied fields up to 10 kOe.
70
Ferroic Materials: Synthesis and Applications
4. Results and Discussion
4.1. Scanning Electron Microscopy: SEM images for the samples with x = 0.1 sintered at different
temperatures are shown in Fig. 1. The figure indicates that the sample sintered at 1100° C consists
of thin hexagonal plates of diameters typically < 500 nm. These particles therefore have diameters
up to the critical single domain size in hexaferrites. The particles for the sample sintered at 1200° C
grew in size, where the typical particle diameter became 0.5 – 1.0 μm. Upon further increasing the
sintering temperature up to 1300° C, the particle diameter grew further up to diameters typically > 1
μm. The samples with higher x values (Fig. 2 and Fig. 3) demonstrated similar tendency of particle
growth with increasing the sintering temperature. Further, increasing the dopant concentration
seems to result in increasing the particle size. The size distribution for the sample with x = 0.2, for
example, is dominated by particles with diameters < 0.5 μm, while the sample with x = 0.3 is
dominated by particles with diameters > 0.5 μm. It is also evident that samples with increasing x
sintered at 1300° C demonstrated uncontrollable growth of particles, leading to the formation of
thin plates with diameters ~ 10 μm. This is indicative that the increased level of Mo – Zn addition to
the hexaferrites prohibits the formation of distinct small particles with well-defined boundaries.
From these results we can conclude that the particles in the samples sintered at temperatures >
1100° C contain multi-domain magnetic structure.
Fig. 1. SEM images of the samples with x = 0.1 sintered at different temperatures.
Solid State Phenomena Vol. 232
71
Fig. 2. SEM images of the samples with x = 0.2 sintered at different temperatures.
Fig. 3. SEM images of the samples with x = 0.3 sintered at different temperatures.
4.2. X-Ray Diffraction: Fig. 4 shows the refined diffraction patterns for all samples sintered at
1100° C. The patterns of all samples indicated the presence of a major BaM hexaferrite phase with
structure consistent with the standard pattern JCPDS: 00-043-0002, and lattice parameters around a
= 5.89 Å and c = 23.2 Å. In addition, α-Fe2O3, BaMoO4 and ZnFe2O4 oxides were detected as minor
72
Ferroic Materials: Synthesis and Applications
phases whose structures were consistent with the standard patterns JCPDS: 00-033-0664, 01-0894570, and 00-022-1012, respectively. These minor phases were detected by other workers in similar
hexaferrites [50, 52]. The intensities of the diffraction peaks corresponding to the Zn-spinel
(ZnFe2O4) and barium molybdenum oxide phases obviously increased with increasing x. The
relative intensity of the main peak of BaMoO4 at about 2θ = 26.7° increased from about 17.5% of
the main BaM peak for the sample with x = 0.1 up to about 52.6% (three times bigger) for the
sample with x = 0.3. This is a clear indication that Mo ions do not enter the BaM lattice, and are
completely incorporated in forming the BaMoO4 phase. Also, we noticed that the α-Fe2O3
disappeared from the sample with x = 0.2, which leads to the conclusion that this sample consists of
three phases with relative molar ratios of:
0.2 (BaMoO4):0.8 (BaFe12O19):0.55 (ZnFe2O4)
These ratios were derived by fixing the Ba, Mo, and Fe molar ratios to the experimental ratios of
1.0:0.2:10.7. The derived molar ratios of the phases indicate that the ratio of Zn (0.55) is slightly
smaller than the starting value of 0.60, which could be attributed to zinc being more volatile than
other elements in the sample. Further, the small amount of the α-Fe2O3 phase in the sample with x =
0.3 is due to the excess of Fe ions which remain in the form of unreacted α-Fe2O3 phase. If the
molar ratios of the starting materials of Ba:Fe:Mo:Zn = 1.0:10.3:0.3:0.9 were used to derive the
molar ratios of the four existing phases, we would obtain:
0.3 (BaMoO4):0.7(BaFe12O19):0.9 (ZnFe2O4):0.05 (α-Fe2O3)
This suggests the presence of the α-Fe2O3 phase with a mass ratio of about 0.7% in the sample with
x = 0.3. Rietveld refinement of the XRD pattern for this sample, however, indicates the presence of
α-Fe2O3 phase with a mass ratio of about 3%. This also could be attributed to Zn-loss in the
sintering process. If we assume 8% loss of Zn (as in the sample with x = 0.2), then the molar ratio
of α-Fe2O3 becomes about 0.24, which accounts for about 3.5% of the mass of the sample, in good
agreement with the experimentally detected relative proportion of this phase.
In an attempt to eliminate the secondary phases and promote the production of a hexaferrite phase
with higher quality, the samples were sintered at higher temperatures of 1200° C and 1300° C,
respectively. It is obvious from Fig. 5 that the Zn-spinel phase disappeared from all samples
sintered at 1200° C, except that with x = 0.3. The presence of the M-type phase as the major phase
and BaMoO4 and α-Fe2O3 as minor phases in the patterns of the samples with x < 0.2 suggest that at
relatively low Zn concentrations (< 0.6), the high-temperature reaction of the Zn-spinel with BaM
promotes the substitution of Zn ions in the BaM lattice, which is evidenced by the change in the
relative intensities of the main peaks for this phase. The appearance of the α-Fe2O3 phase in these
samples is facilitated by the formation of BaMoO4 phase and the consequent reduction in the
amount of Ba which is required to react with α-Fe2O3 to form the hexaferrite phase. On the other
hand, the persistence of the Zn-spinel phase, the variation of the relative intensities of the BaM
peaks, and the observed increase of the α-Fe2O3 phase in the sample with x = 0.3 compared to the
sample sintered at 1100° C (Fig. 6) is an indication of the limited solubility of Zn in the BaM
hexaferrite lattice. In the sample with x = 0.2, however, W-type hexaferrite phase whose structure is
consistent with the standard pattern JCPDS: 00-052-1868 for Zn2W (BaZn2Fe16O27) developed, and
the Zn-spinel phase disappeared almost completely. This result is evidence that the reaction of Znspinel and BaM phases is completed to form Zn2W at 1200° C in this sample in accordance with the
reaction [52]:
2 (ZnFe2O4) + BaFe12O19 = BaZn2Fe16O27
(2)
Solid State Phenomena Vol. 232
73
Accordingly, the relative intensities of the diffraction peaks corresponding to the W-phase increased
at the expense of those corresponding to BaM phase (Fig. 7). It should be kept in mind, however,
that the W-phase can exist with the general chemical formula BaZn2-yFe16+yO27, where the value of
y is determined by the amount Zn ions substituting Fe ions in the BaM lattice, and has a significant
effect on the magnetic properties of the compound [1].
Fig. 4. XRD patterns of the BaFe12-4xMoxZn3xO19 samples sintered at 1100° C.
74
Ferroic Materials: Synthesis and Applications
Fig. 5. XRD patterns of the BaFe12-4xMoxZn3xO19 samples sintered at 1200° C.
Solid State Phenomena Vol. 232
Fig. 6. XRD patterns of BaFe10.8Mo0.3Zn0.9O19 samples sintered at different temperatures.
75
76
Ferroic Materials: Synthesis and Applications
Fig. 7. XRD patterns of BaFe11.2Mo0.2Zn0.6O19 samples sintered at different temperatures.
Upon sintering the samples at 1300° C, the XRD patterns (Fig. 8) indicate that the Zn-spinel phase
disappeared completely from all samples, and the W-phase appeared in the sample with x = 0.3 as a
consequence of the reaction of the Zn-spinel and BaM intermediate phases. However, for x < 0.2
the W-type (BaZn2Fe16O27) phase was not observed, which is an indication of the limited solubility
of Zn ions in the BaM phase, and that the critical concentration of Zn ions which is required to
Solid State Phenomena Vol. 232
77
produce the W-phase is 0.6. Accordingly, in the subsequent sections we will limit our hyperfine and
magnetic studies to samples with x = 0.2 and 0.3.
Fig. 8. XRD patterns of the BaFe12-4xMoxZn3xO19 samples sintered at 1300° C.
4.3. Mössbauer Spectroscopy: Mössbauer spectroscopy (MS) is based on a local effect which can
provide structural as well as magnetic information [53 – 55]. This effect uses 57Fe as a probe. Iron
ions residing at different sites with different coordination and chemical environments give rise to
different components in Mössbauer spectrum of a sample. Magnetically ordered iron ions in a given
sublattice give rise to a six-line pattern (sextet) magnetic component due to Zeeman splitting.
However, a magnetically disordered Fe sublattice gives a paramagnetic component: a singlet or a
doublet. The doublet typically has a quadrupole splitting which is dependent on the iron valence
state, and the distortion of the iron site and deviation from cubic symmetry. This quadrupole
splitting arises from the interaction of the nuclear moment of 57Fe with the electric field gradient at
the nuclear site.
Fig. 9 shows Mössbauer spectrum for the sample with x = 0 sintered at 1100° C. The spectrum was
fitted with five magnetic components corresponding to the five iron sites in the lattice (Table 1).
The fitting parameters (Table 3) are in very good agreement with previously reported results [28,
51, 56]. The values of isomer shifts (~ 0.3 – 0.4 mm/s) are typical for Fe3+. The results of MS
confirm the structural analysis which indicated the presence of a single BaM phase in this sample.
78
Ferroic Materials: Synthesis and Applications
Intensity
x=0
-15
-10
-5
0
5
10
15
Velocity (mm/s)
Fig. 9. Mössbauer spectrum for BaFe12O19 sintered at 1100° C.
Table 3. The hyperfine fields (Bhf) in kOe, isomer shifts (CS) in mm/s, quadrupole splitting (QQ) in
mm/s, and percentage relative intensities (I) of the components of the spectra for the samples with
different values of x sintered at the indicated temperature.
Mössbauer
parameters
Msites
x = 0.0
x = 0.2
x = 0.2
W-sites
x = 0.3
(1100° C)
(1100° C)
(1200° C)
Bhf1
4f2
518
518
516
4fVI
Bhf2
2a
511
504
494
6g
Bhf3
4f1
494
493
479
4f
486
Bhf4
12k
418
418
421
4e+4fIV
418
Bhf5
2b
405
411
383
2d
386
Bhf6
-
-
-
373
12k
369
CS1
0.39
0.40
0.40
CS2
0.38
0.43
0.27
CS3
0.27
0.28
0.41
0.32
CS4
0.38
0.38
0.38
0.38
CS5
0.32
0.33
0.29
0.43
(1100° C)
511
0.42
Solid State Phenomena Vol. 232
79
CS6
-
0.38
0.37
0.38
QQ1
0.23
0.19
0.19
0.12
QQ2
0.12
0.10
0.28
QQ3
0.22
0.21
0.16
0.18
QQ4
0.42
0.41
0.39
0.40
QQ5
2.19
2.24
2.24
2.50
-
0.34
0.34
0.39
I1
19.7
11.8
9.8
18.0
I2
6.8
10.9
15.9
24.7
I3
16.4
14.2
10.1
I4
50.8
47.0
25.9
27.8
I5
6.3
5.4
5.8
2.8
10.8
32.4
26.7
2.70
2.22
14.2
QQ6
I6
χ2
-
2.68
Fig. 10 shows Mössbauer spectrum for the sample with x = 0.2 sintered at 1100° C and 1200° C.
Evidently, the magnetic sub-spectrum for the sample sintered at 1100° C is similar to the spectrum
for the sample with x = 0.0. In addition, a significant paramagnetic component appeared near the
center of the spectrum. Accordingly, the spectrum was fitted with five magnetic components and
one paramagnetic doublet. The hyperfine parameters for the magnetic components are similar to
those for the sample with x = 0, and these components are therefore associated with the five iron
sites in BaM lattice. The paramagnetic doublet is associated with the paramagnetic ZnFe2O4 spinel
phase [57] which was confirmed by our XRD analyses.
The spectrum for the sample with x = 0.2 sintered at 1200° C, however, shows a significantly
different structure from that of the sample sintered at 1100° C. In addition, the central
paramagnetic component associated with the zinc-spinel phase disappeared at this sintering
temperature, which is consistent with the XRD data. The spectrum was best fitted with six magnetic
components with hyperfine parameters listed in Table 3 above. The hyperfine parameters for the
magnetic components are consistent with those of previously reported results on other W-type
hexaferrites [57, 58]. This result is also consistent with the presence of BaZn2Fe16O27 phase
observed in XRD pattern for this sample. In this phase there are seven distinct crystallographic
sites: two tetrahedral (4e and 4fIV), four octahedral (4fVI, 6g, 4f, and 12k) and one bi-pyramidal (2d)
sites (see Table 2). The hyperfine parameters for the 4e and 4fIv cannot be distinguished from each
other, and one magnetic component is typically assigned to these two sites. In some cases when the
spectrum is complex, the magnetic components associated with 6g and 4f sites also cannot be
distinguished from each other. The six-component fit for the sample with x = 0.2 sintered at 1200°
C gave relative sub-spectral intensities which are in good agreement with the theoretical relative
ratios of 2:3:2:4:1:6.
80
Ferroic Materials: Synthesis and Applications
Intensity
XRD results indicated the coexistence of the Fe-containing BaM and Zn2W phases in this sample.
However, due to the similarity of the crystal structures and symmetries of the sites for these two
hexaferrites, and the complexity of the Mössbauer spectrum, it was not possible to resolve the
components for the two different types. However, the persistence of the strong, sharp component
associated with the 12k sublattice of BaM could be evidence of the presence of the two phases in
this sample.
Intensity
x=0.2, 1100C
x=0.2, 1200C
-15
-10
-5
0
5
10
15
Velocity (mm/s)
Fig. 10. Mössbauer spectrum for BaFe11.2 Mo0.2Zn0.6O19 sintered at 1100 and 1200° C.
Mössbauer spectrum for the sample with x = 0.3 sintered at 1100° C is shown in Fig. 11. The
spectrum for this sample shows a magnetic sextet and a central paramagnetic component, similar to
the spectrum for the sample with x = 0.2 sintered at 1100º C. The spectrum was best fitted with five
magnetic components and a paramagnetic component. The hyperfine parameters of the magnetic
sextets (Table 3) are consistent with the five different sites of BaM. The two magnetic components
corresponding to the 2a and 4f1 sites, however, could not be resolved, and appeared as a single
component [16]. The relative intensity (24.7%) of this magnetic sub-spectrum is in good agreement
with the theoretical value of 25%. In addition, a new component with relatively low hyperfine field
(369 kOe) developed, which is associated with the splitting of the 12k component as a consequence
of the perturbation of part of the corresponding sublattice by excessive substitution of nonmagnetic
ions at neighboring sites [17]. The reduction of the hyperfine fields of the sextets is an indication of
the perturbation of the magnetic sublattices by the substitution process. In the context of discussing
the structural properties of this sample, it was mentioned that the Zn-spinel phase does not account
for the amount of Zn in the precursor powder, and this was associated with the evaporation of Zn
Solid State Phenomena Vol. 232
81
during the sintering process. On the basis of the Mössbauer data, however, we could argue that
some Zn ions entered the hexaferrite lattice, resulting in the observed variations of the hyperfine
parameters. The central paramagnetic component is associated with the Zn-ferrite phase confirmed
by the XRD pattern of this sample.
Intensity
x=0.3
-15
-10
-5
0
5
10
15
Velocity (mm/s)
Fig. 11. Mössbauer spectrum for (BaFe11.8 Mo0.3Zn0.9O19) sintered at 1100° C
4.4. Magnetic Measurements: The magnetic properties of the hexaferrites prepared with x = 0.2
and 0.3 are addressed in the following discussion. The hysteresis properties are discussed in light of
the structural characteristics and the effects of heat treatment is discussed. Further, the
thermomagnetic measurements were used to shed light on the magnetic phases and the magnetic
transition temperatures in the samples.
4.4.1. Hysteresis Characteristics: The hysteresis loops for the samples with x = 0.0, 0.2, and 0.3
sintered at 1100° C are shown in Fig. 12. The magnetization did not saturate up to the maximum
field applied, and therefore, the saturation magnetization was determined from the law of approach
to saturation, according to which the magnetization in the high field region is given by [59]:
𝑀 = 𝑀𝑠 (1 −
𝐴
𝐵
− 2 ) + 𝜒𝐻
𝐻 𝐻
(3)
Here M is the magnetization (in emu/cm3), A is a constant associated with microstress and/or
inclusions, B is a constant due to magnetocrystalline anisotropy, and the last term in eq. 3 is the
forced magnetization term. A plot of M vs 1/H2 in the high-field region gave a straight line for each
sample, indicating that the contributions of the microstress/ inclusions, and the forced magnetization
are negligible due to the high magnetocrystalline anisotropy in our samples. The saturation
magnetization σs (= Ms divided by the density) was obtained from the intercept of the straight line,
and the results are tabulated in Table 4. The decrease in magnetization with increasing x in the
samples sintered at 1100° C is mainly due to the increase in the relative proportions of the
nonmagnetic phases as confirmed by XRD.
82
Ferroic Materials: Synthesis and Applications
The coercivity and remnant magnetization for each sample were determined from the hysteresis
loop, and the results are tabulated in Table 4. We observed that the coercivity decreased with
increasing x from a hard magnetic state ( HC = 3270 Oe for x = 0.0) to a state of moderate magnetic
hardness ( HC = 910 Oe for x = 0.3). Although the magnetic phase in all samples is BaM, the
modification of the coercivity could be associated with the different particle size distributions in the
samples, and the growth of the mean particle diameter with increasing x [58]. SEM image of the
sample with x = 0.3 indicated the presence of large multi-domain particles which should have a
significant effect on lowering the coercivity of this sample with respect to that with x = 0.2, whose
SEM image demonstrated that the particle size is dominated by single domain particles. Further, the
squareness ratio (= σr/ σs) for the samples with x = 0.0 and 0.2 is ~ 0.5, which is characteristic of
randomly oriented single-domain particles (with typical diameters < 0.5 µm). However, the
squareness ratio for the sample with x = 0.3 is ~ 0.4, which is a further confirmation that this sample
consists of multi-domain particles, and thus has lower coercivity due to the dominance of domainwall motion in the magnetization processes.
60
40
T = 1100 C
x=0
x = 0.2
x = 0.3
 (emu/g)
20
0
-20
-40
-60
-10000
-5000
0
5000
10000
H (Oe)
Fig. 12. Hysteresis loops of BaFe12-4xMoxZn3xO19 ferrites sintered at 1100° C.
Solid State Phenomena Vol. 232
83
Table 4. Saturation magnetization σs (emu/g), remnant magnetization σr (emu/g) and coercivity HC
(Oe) of ferrites samples sintered at 1100° C, 1200° C and 1300° C.
x
0
0.2
0.3
1100
σs
67.1
60.1
54.6
σr
35.7
29.6
22.2
Hc
3275
2100
910
1200
σs
71.7
-
σr
20.0
-
1300
Hc
σs
1050 74.4
66.7
σr
14.6
6.4
Hc
447
180
The hysteresis loops for the sample with x = 0.2 sintered at 1200° C and 1300° C are presented in
Fig. 13, and the magnetic parameters derived from these loops are listed in Table 4. It is evident that
increasing the sintering temperature improved the saturation magnetization, and reduced the
coercivity significantly. The sample sintered at 1200° C exhibited a saturation magnetization of
71.7 emu/g and coercivity of 1050 Oe. While the observed saturation magnetization is in very good
agreement with that (71 emu/g) reported for multi-domain Fe2W particles sintered at 1200° C, the
coercivity is significantly higher than the reported value of 305 Oe for Fe2W particles [58]. This
difference could be due to the coexistence of the hard M-type phase in our sample, although effects
of different particle size and chemical stoichiometry cannot be excluded. As the sintering
temperature increased up to 1300° C, the saturation magnetization increased up to 74.4 emu/g, and
the coercivity dropped down to 447 Oe. These values are in excellent agreement with those of Fe2W
particles sintered at 1250° C [58] and slightly lower than the saturation magnetization (78 emu/g) of
bulk Fe2 hexaferrite [1]. The increase in saturation magnetization upon sintering at 1300° C is
associated with the increased level of transformation of the BaM phase to W-type phase, and the
observed saturation magnetization (with value intermediate between that of the M-type and that of
the W-type) could indicate that the two phases coexist with almost equal mass fractions. The
secondary barium molybdenum oxide in the sample also should have an effect in reducing the
saturation magnetization, but due to its small mass fraction in the sample, its effect appears to be
negligible, and the obtained samples could be considered high quality hexaferrite samples with
tunable magnetic properties.
80
60
40
x = 0.2
T = 1200 C
T = 1300 C
(emu/g)
20
0
-20
-40
-60
-80
-10000
-5000
0
5000
10000
H (Oe)
Fig. 13. Hysteresis loops of BaFe11.2Mo0.2Zn0.6O19 ferrites sintered at 1200° C and 1300° C.
84
Ferroic Materials: Synthesis and Applications
As the optimal sintering temperature for the formation of the W-type phase in the samples with x =
0.2 and 0.3 seems to be 1300° C, the hysteresis loops of these two samples were measured and
analyzed. Fig. 14 and the data in Table 4 indicate that the saturation magnetization of the sample
with x = 0.3 is lower than that of the sample with x = 0.2. This marginal decrease could be
associated with the higher presence of nonmagnetic barium molybdenum oxide as evidenced by the
XRD patterns of the two samples. The coercivity is further reduced with increasing x, which could
be due to the increased particle diameter. The produced compounds could be suitable for
applications requiring soft magnetic materials.
80
60
T = 1300 C
x = 0.2
x = 0.3
40
(emu/g)
20
0
-20
-40
-60
-80
-10000
-5000
0
5000
10000
H (Oe)
Fig. 14. Hysteresis loops of BaFe12-4xMoxZn3xO19 ferrites sintered at 1300° C.
4.4.2. Initial Magnetization Curves: The magnetization curves for the samples presented in Fig.
15 are clearly different in more than one aspect. The sample with x = 0.2 sintered at 1100° C has
lower magnetization than the un-doped BaM sample as mentioned in the above discussion.
However, this sample is clearly softer, as its initial susceptibility (σ/H) is about 35% higher than
that of BaM. This result is consistent with the growth of the particles upon substitution for Fe ions
in this system. At higher sintering temperatures, the samples became progressively softer, and the
magnetization increased. The initial susceptibility of the sample with x = 0.2 sintered at 1200° C
increased by up to 3.6 times that of BaM, which is also consistent with the previously discussed
structural and magnetic properties of this sample. Further, the sample with x = 0.3 sintered at 1300°
C is clearly a soft magnetic material with an initial susceptibility ~ 6.9 times greater than that of
BaM. The largest increase in susceptibility is induced by the highest degree of transformation of
BaM to W-type, and the increase in particle size.
Solid State Phenomena Vol. 232
85
70
60
(emu/g)
50
40
30
x = 0.0, T = 1100 C
x = 0.2; T = 1100 C
x = 0.2; T = 1200 C
x = 0.3; T = 1300 C
20
10
0
0
2000
4000
6000
8000
10000
H (Oe)
Fig. 15. Magnetization curves of BaFe12-4xMoxZn3xO19 ferrites sintered at the indicated
temperatures.
4.4.3. Thermomagnetic Measurements: The thermomagnetic curve of the sample with x = 0
measured at an applied field of 8 kOe is shown in Fig. 16, together with its derivative with respect
to temperature. The derivative shows a single dip at 450 °C, indicating that this sample consists of a
single BaM phase.
60

Derivative
50
0.00
H = 8 kOe
-0.05
(emu/g)
-0.15
30
-0.20
20
-0.25
Derivative
-0.10
40
-0.30
10
-0.35
0
0
100
200
300
400
500
-0.40
600
T (deg. C)
Fig. 16. Thermomagnetic curve of BaFe12O19 ferrites sintered at 1100° C.
86
Ferroic Materials: Synthesis and Applications
The thermomagnetic curve of the sample with x = 0.2 sintered at 1100° C (Fig. 17) shows a dip in
the derivative at about 455° C, which is associated with the Curie temperature of BaM phase. The
closeness of the transition temperature to that of pure BaM may indicate that substitution did not
take place in this sample. In addition, a clear dip is observed at about 225° C. This transition
temperature could be associated with (ZnxFe1-x)Fe2O4 spinel ferrite phase observed in XRD. The
peak at about 115° C, however, could be due to spin-glass transition at this temperature. The shape
of the thermomagnetic curve is, however, dominated by BaM behavior, which is consistent with our
XRD and Mössbauer studies.

(emu/g)
Derivative
0.00
H = 8 kOe
40
-0.05
30
-0.10
20
-0.15
10
-0.20
0
-0.25
0
100
200
300
400
500
Derivative
50
600
T (deg. C)
Fig. 17. Thermomagnetic curve of BaFe11.2 Mo0.2Zn0.6O19 ferrite sintered at 1100° C.
The thermomagnetic curve of the sample with x = 0.2 sintered at 1200° C shows far more rich
structure as indicated by Fig. 18. The dip at 410° C is associated with the Curie temperature of the
BaM phase. The decrease in the transition temperature (by about 40° C) could be associated with
partial substitution of Zn ions at the tetrahedral site of the BaM lattice, resulting in a decrease in the
strength of the super-exchange interactions between spin-up and spin-down sublattices. The dip at
350° C is associated with the transition temperature of the W-phase. The fact that this transition
temperature is significantly lower than that of ZnFe-W phase (450° C [1]) may indicate that the
majority of the Me2+ ions in the W-phase in our sample are Zn2+. This low transition temperature is,
however, consistent with the 300° C – 400° C reported for substituted Co2W hexaferrite [60]. The
clear dip at 217° C is associated with the Zn-spinel ferrite, part of which may have remained at this
temperature. Due to the complexity of the magnetic and crystallographic structure of this sample,
however, we cannot confirm such an assignment of the transition temperature. In light of our
structural analysis, which confirmed the presence of BaM and W-type phases, an alternative
assignment can be suggested, according to which the transition temperature at 217° C is associated
with a secondary W-type phase which is rich in Zn and has lower transition temperature. In a yet
second alternative argument, one could suggest that the dips at 410° C and 350° C are associated
with the Curie temperatures of the BaM phases with different levels of Zn substitution. These
transition temperatures are consistent with reported temperatures in the range 292 – 402° C for
substituted BaM hexaferrites [61]. The transition temperature at 217° C is associated with the Znrich (Zn,Fe)-W phase. These arguments seem to be more consistent with the XRD and Mössbauer
data.
Solid State Phenomena Vol. 232
87
0.05
60

H = 8 kOe
Derivative
0.00
50
-0.05
-0.10
30
-0.15
20
Derivative
(emu/g)
40
-0.20
10
-0.25
0
0
100
200
300
400
500
-0.30
600
T (deg. C)
Fig. 18. Thermomagnetic curve of BaFe11.2 Mo0.2Zn0.6O19 ferrite sintered at 1200° C.
The thermomagnetic curve of the sample with x = 0.3 sintered at 1300° C (Fig. 19) clearly indicates
two major transitions at 344° C and 225° C. In light of the XRD data of this sample, which indicates
the coexistence of BaM and W-type hexaferrite phases, these transition temperatures are associated
with the two hexaferrite phases, and seem to support the second alternative argument in the
discussion above.
0.05
(emu/g)
Derivative
H = 8 kOe
0.00
50
-0.05
40
-0.10
-0.15
30
-0.20
20
-0.25
10
-0.30
Derivative

60
-0.35
0
0
100
200
300
400
500
-0.40
600
T (deg. C)
Fig. 19. Thermomagnetic curve of BaFe11.2 Mo0.3Zn0.9O19 ferrite sintered at 1300° C.
88
Ferroic Materials: Synthesis and Applications
5. Conclusions
Rich structural and magnetic phase diagram was found in Mo–Zn substituted barium hexaferrites
with the stoichiometry of BaM phase. The samples were prepared by ball milling and subsequent
sintering at temperatures from 1100° C to 1300° C. The samples sintered at 1100° C consisted of a
major M-type phase in addition to barium-molybdenum oxide, Zn-spinel, and α-Fe2O3 oxides. The
saturation magnetization decreased with the level of substitution due to the increased relative
fractions of the nonmagnetic phases. However, the saturation magnetization of the substituted
samples remained relatively high (> 54 emu/g). Also, coercivity values of about 2100 and 900 Oe
were obtained by substitution with x = 0.2 and 0.3, respectively. Upon increasing the sintering
temperature, the W-phase evolved as a major phase, the saturation magnetization improved, and the
samples transformed to soft magnetic materials at 1300° C sintering temperature. Our study
indicated that the sample with x = 0.2 possesses the best magnetic properties for magnetic recording
and soft magnetic materials applications, where the magnetic properties can be tuned by appropriate
heat treatment in a wide range of coercivity, while maintaining a high saturation magnetization.
Acknowledgements
This work was supported by a generous fund from the Deanship of Scientific Research at The
University of Jordan (Grant # 1404). The technical assistance of Yousef Abu Salha and Waddah
Fares (The University of Jordan) is acknowledged.
References
[1]
J. Smit and H.P.J. Wijn, Ferrites, (Philips Technical Library, Eindhoven, 1959).
[2]
Ferromagnetic Materials, Vol. 2, E.P. Wohlfarth, ed. (North-Holland, Amsterdam, 1980).
[3]
R.C. Pullar, Hexagonal ferrites: A review of the synthesis, properties and applications of
hexaferrite ceramics, Progress in Materials Science 57 (2012) 1191 – 1334.
[4]
U. Ozgur, Y. Alivov, and H. Morkoc, Microwave Ferrites, Part 1: Fundamental Properties,
Journal of Materials Science: Materials in Electronics 20 (2009) 789 – 834.
[5]
G. Albanese, Recent advances in hexagonal ferrites by the use of nuclear spectroscopic
methods, Journal de Physique 38 (1977) C1-85 – C1-94.
[6]
N. Dishovski, A. Petkov, Iv. Nedkov, and Razkazov, Hexaferrite contribution to microwave
absorbers characteristics, IEEE Transactions on Magnetics 30 (1994) 969 – 971.
[7]
R.S. Meena, S. Bhattachrya, and R. Chatterjee, Coplex permittivity, permeability and
microwave absorbing properties of (Mn2-xZnx)U-type hexaferrite, Journal of Magnetism and
Magnetic Materials 322 (2010) 2908 – 2914.
[8]
S.H. Mahmood, F.S. Jaradat, A.-F. Lehlooh, and A. Hammoudeh, Structural properties and
hyperfine interactions in Co–Zn Y-type hexaferrites prepared by sol–gel method, Ceramics
International 40 (2014) 5231 – 5236.
[9]
V.G. Harris, A. Geiler, Y. Chen, S.D. Yoon, M. Wu, A. Yang, Z. Chen, P. He, P.V. Parimi,
X. Zuo, C.E. Patton, M. Abe, O. Acher, and C. Vittoria, Journal of Magnetism and Magnetic
Materials 321 (2009) 3035 – 3047.
[10] T. Tanaka, T. Jitosho, and K. Yamamori, Overwrite and bit shift characteristics of Ba-ferrite
floppy disks, Journal of Magnetism and Magnetic Materials 134 (1994) 390 – 394.
Solid State Phenomena Vol. 232
89
[11] D.E. Speliotis, Advanced MP++ and BaFe++ tapes, Journal of Magnetism and Magnetic
Materials 155 (1996) 83 – 85.
[12] Z. Yang, C.S. Wang, X.H. Li, and H.X. Zeng, (Zn, Ni, Ti) substituted barium ferrite particles
with improved temperature coefficient of coercivity, Materials Science and Engineering B90
(2002) 142 – 145.
[13] I.V. Zavislyak, M.A. Popov, and G. Srinivasan, A cut-off millimeter wave resonator
technique for mapping magnetic parameters in hexagonal ferrites, Measurement Science and
Technology 20 (2009) 115704 (5 pp).
[14] G. Bate, in Ferromagnetic Materials, Vol. 2, E.P. Wohlfarth, ed. (North-Holland, Amsterdam,
1980) 381 – 507.
[15] A. González-Angeles, J. Lipka, A. Grusková, V. Jančárik, I. Tóth, and J. Sláma, (Ni, Zn, Sn)
Ru and (Ni, Sn) Sn substituted barium ferrite prepared by mechanical alloying, Hyperfine
Interactions 184 (2008) 135 – 141.
[16] I. Bsoul, S. H. Mahmood, and A.-F. Lehlooh, Structural and magnetic properties of
BaFe12−2xTixRuxO19, Journal of Alloys and Compounds 489 (2010) 157–161.
[17] I. Bsoul, S.H. Mahmood, A-F. Lehlooh, and A. Al-Jamel, Structural and magnetic properties
of SrFe12-2xTixRuxO19, Journal of Alloys and compounds 551 (2013) 490 – 495.
[18] S. Choopani, N. Keyhan, A. Ghasemi, A. Sharbati, and R.S. Alam, Structural, magnetic and
microwave absorption characteristics of BaCoxMnxTi2xFe12-4xO19, Materials Chemistry and
Physics 113 (2009) 717 – 720.
[19] G.B. Teh, N. Saravanan, and D.A. Jefferson, A study of magnetoplumbite-type (M-type)
cobalt-titanium substituted barium ferrite, BaCoxTixFe12-2xO19 (x = 1 – 6), Materials Chemistry
and Physics 105 (2007) 253 – 259.
[20] G. Mendoza-Suárez, L.P. Rivas-Vázquez, J.C. Corral-Huacuz, A.F. Fuentes, and J.I.
Escalante-García, Magnetic properties and microstructure of BaFe11.6-2xTixMxO19 (M = Co,
Zn, Sn) compounds, Physica B 339 (2003) 110 – 118.
[21] S. Nilpairach and W. Udomkichdaecha, Coercivity of the co-precipitated prepared
hexaferrites, BaFe12-2xCoxSnxO19, Journal of the Korean Physical Society 48 (2006) 939 –
945.
[22] A. Ghasemi, A. Hossienpour, A. Morisako, A. Saatchi, and M. Salehi, Electromagnetic
properties and microwave absorbing characteristics of doped barium hexaferrites, Journal of
Magnetism and Magnetic Materials 302 (2006) 429 – 435.
[23] A. González-Angeles, G. Mendoza-Suárez, A. Grusková, J. Sláma, J. Lipka, and M.
Papánová, Magnetic structure of Sn2+Ru4+ -substituted barium hexaferrites prepared by
mechanical alloying, Materials letters 59 (2005) 1815_1819.
[24] V.V. Soman, V.M. Nanoti, and D.K. Kulkarni, Dielectric and magnetic properties of Mg–Ti
substituted barium hexaferrites, Ceramics International 39 (2013) 5713 – 5723.
[25] Y. Lui, M.G.B. Drew, and Ying Liu, Preparation and magnetic properties of barium ferrites
substituted with manganese, cobalt, and tin, Journal of Magnetism and Magnetic Materials
323 (2011) 945 – 953.
[26] A.M. Alsmadi, I. Bsoul, S.H. Mahmood, G. Alnawashi, K. Prokeš, K. Siemensmeyer, B.
Klemke, and H. Nakotte, Magnetic study of M-type doped barium hexaferrite nanocrystalline
particles, Journal of Applied Physics 114 (2013) 243910 (8 pp).
90
Ferroic Materials: Synthesis and Applications
[27] M. Manawan, A. Manaf, B. Soegijono, and A. Yudi, Microstructural and magnetic properties
of Ti2+-Mn4+ substituted barium hexaferrite, Advanced Materials Research, 896 (2014) 401 –
405.
[28] M. Awawdeh, I. Bsoul, and S.H. Mahmood, Magnetic properties and Mössbauer spectroscopy
on Ga, Al, and Cr substituted hexaferrites, Journal of Alloys and Compounds 585 (2014) 465
– 473.
[29] S. Ounnunkad and P. Winotai, Properties of Cr-substituted M-type barium ferrites prepared
by nitrate–citrate gel-autocombustion process, Journal of Magnetism and Magnetic Materials
301 (2006) 300–301.
[30] I. Bsoul and S.H. Mahmood, Magnetic and structural properties of BaFe12- xGaxO19
nanoparticles, Journal of Alloys and Compounds 489 (2010) 110 – 114.
[31] D. Chen, Y. Liu, Y. Li, K. Yang, and H, Zhang, Microwave and magnetic properties of Aldoped barium ferrite with sodium citrate as chelate agent, Journal of magnetism and magnetic
Materials 337 – 338 (2013) 65 – 69.
[32] P. Sharma, R.A. Rocha, S.N de Medeiros, and A. Paesano Jr, Structural and magnetic studies
on barium hexaferrites prepared by mechanical alloying and conventional route, Journal of
Alloys and Compounds 443 (2007) 37 – 42.
[33] S.R. Janasi, D. Rodrigues, F.J.G. Landgraf, and M. Emura, Magnetic properties of
coprecipitated barium ferrite powders as a function of synthesis conditions, IEEE
Transactions on Magnetics 36 (2000) 3327 – 3329.
[34] S.R. Janasi, M. Emura, F.J.G. Landgraf, and D. Rodrigues, The effects of synthesis variables
on the magnetic properties of coprecipitated barium ferrite powders, Journal of Magnetism
and Magnetic Materials 238 (2002) 168–172.
[35] Y.Y. Meng, M.H. He, Q. Zeng, D.L. Jiao, S. Shukla, R.V. Ramanujan, and Z.W. Liu,
Synthesis of barium ferrite ultrafine powders by sol–gel combustion method using glycine
gels, Journal of Alloys and Compounds 583 (2014) 220 – 225.
[36] U. Topal, H. Ozkan, and K.G. Topal, Improved properties of BaFe12O19 prepared by
ammonium nitrate melt technique and washed in HCl, Journal of Alloys and Compounds 422
(2006) 276–278.
[37] H. Sӧzeri, Z. Durmuş, A. Baykal, and E. Uysal, Preparation of high quality, single domain
BaFe12O19 particles by the citrate sol–gel combustion route with an initial Fe/Ba molar ratio
of 4, l Materials Science and Engineering B 177 (2012) 949 – 955.
[38] M. Han, Y. Ou, W. Chen, and L. Deng, Magnetic properties of Ba-M-type hexagonal ferrites
prepared by the sol–gel method with and without polyethelene glycol added, Journal of
Alloys and compounds 474 (2009) 185 – 189.
[39] X. Tang, B.Y. Zhao, and K.A. Hu, Preparation of M-Ba-ferrite fine powders by sugar nitrates
processes, Journal of Materials Science 41 (2006) 3867 – 3871.
[40] J. Qiu, L. Lan, H. Zhang, and M. Gu, Effect of titanium dioxide on microwave absorption
properties of barium ferrite, Journal of Alloys and Compounds 453 (2008) 261 – 264.
[41] J. Lipka, A. Grusková, O. Orlicky, J. Sitek, M. Miglierini, R. Grone, M. Hucl, and I. Tóth,
Mössbauer and magnetic susceptibility measurements on M-type hexagonal Ba-ferrite,
Hyperfine Interactions 59 (1990) 381 – 386.
Solid State Phenomena Vol. 232
91
[42] E. Kreber and U. Gonser, Determination of cation distribution in Ti4+ and Co2+ substituted
barium ferrite by Mössbauer spectroscopy, Applied Physics 10 (1976) 175 – 180.
[43] S.E. Jacobo, C. Domingo-Pascual, R. Rodriguez-Clemente, and M.A. Blesa, Synthesis of
ultrafine particles of barium ferrite by chemical coprecipitation, Journal of Materials Science
32 (1997) 1025 – 1028.
[44] A. Grusková, J. Sláma, R. Dosoudil, D. Kevická, V. Jančárik, and I. Tóth, Influence of Co –
Ti substitution on coercivity in Ba ferrite, Journal of magnetism and Magnetic Materials 242245 (2002) 423 – 425.
[45] Y.L. Chen, X.D. Li, and B.F. Xu, A Mössbauer study of the bypyramidal lattice site (2b) in
BaFe12O19, Hyperfine Interactions 62 (1990) 219 – 224.
[46] F. Gao, D. Li, and S. Zhang, Mössbauer spectroscopy and chemical bonds in BaFe12O19
hexaferrites, Journal of Physics: Condensed Matter 15 (2003) 5079 – 5084.
[47] T.M. Meaz and C. Bender Koch, A crystallographic and Mössbauer spectroscopic study of
BaCo0.5xZn0.5xTixFe12-2xO19 (M-type hexagonal ferrite), Hyperfine Interactions 156/157 (2004)
341 – 346.
[48] S. Chikazumi, Physics of Ferromagnetism, (Oxford University Press, Oxford, 1997).
[49] D.G. Agresti, T.D. Shelfer, A Mössbauer stuy of CoMo-substituted barium ferrite, IEEE
Tranactions on Magnetics 25 (1989) 4069 – 4071.
[50] G.H. Dushaq, S.H. Mahmood, I. Bsoul, H.K. Juwhari, B. Lahlouh, and M.A. AlDamen,
Effects of Molybdenum concentration and valence state effects on the structural and magnetic
properties of BaFe11.6MoxZn0.4-xO19 hexaferrites, Acta Metallurgica Sinica (English letters) 26
(2013) 509 – 516.
[51] S.H. Mahmood, G.H. Dushaq, I. Bsoul, M. Awawdeh, H.K. Juwhari, B. Lahlouh, and M.A.
AlDamen, Magnetic properties and hyperfine interactions in M-type BaFe12-2xMoxZnxO19
hexaferrites, Journal of Applied Mathematics and Physics 2 (2014) 77 – 87.
[52] S.H. Mahmood, A.N. Aloqaily, Y. Maswadeh, A. Awadallah, I. Bsoul, and H. Juwhari,
Structural and magnetic properties of Mo-Zn substituted (BaFe12-4xMoxZn3xO19) M-type
hexaferrites, Material Science Research (India) 11 (2014) 9 – 20.
[53] Mössbauer Spectroscopy, D.P.E Dickson and F.J. Berry, eds. (Cambridge University Press,
Cambridge, 1986).
[54] Mössbauer Spectroscopy, U. Gonser, ed. (Springer-Verlag, New York, 1975).
[55] N.N. Greenwood and T.C. Gibb, Mössbauer Spectroscopy, (Chapman and Hall Ltd, London,
1971).
[56] G.K. Thompson and B.J. Evans. The structure-property relationship in M-type hexaferrites:
Hyperfine Interactions and bulk magnetic properties, Journal of Applied Physics 73 (1993)
6295 – 6297.
[57] G. Albanese, J.L. Sanchez, G. Lopez, S. Diaz, F. Leccabue, and B.E. Watts, Mössbauer and
magnetic study of Sr(ZnLi0.5Fe0.5)Fe16O27 hexaferrite, Journal of Magnetism and Magnetic
Materials 137 (1994) 313 – 321.
[58] H. Kojima, C. Miyakawa, T. Sato, and K. Goto, Magnetic properties of W-type hexaferrite
powders, Journal of Applied Physics 24 (1985) 51 – 56.
92
Ferroic Materials: Synthesis and Applications
[59] B.D. Cullity and C.D. Graham, Introduction to Magnetic Materials, 2nd ed. (Wiley, Hoboken,
NJ, 2009).
[60] M.J. Iqbal, R.A. Khan, S. Mizukami, and T. Miyazaki, Tailoring of structural, electrical and
magnetic properties of BaCo2 W-type hexaferrites by doping with Zr–Mn binary mixtures for
useful applications, Journal of Magnetism and Magnetic Materials 323 (2011) 2137 – 2144.
[61] S.Y. An, I-B. Shim, and C.S. Kim, Mössbauer and magnetic properties of Co-Ti substituted
barium hexaferrite nanoparticles, Journal of Applied Physics 91 (2002) 8465 – 8467.