You are on page 1of 10

Physica B 407 (2012) 795–804

Contents lists available at SciVerse ScienceDirect

Physica B
journal homepage: www.elsevier.com/locate/physb

Infrared and structural studies of Mg1–xZnxFe2O4 ferrites


K.A. Mohammed a,n, A.D. Al-Rawas b, A.M. Gismelseed b, A. Sellai b, H.M. Widatallah b, A. Yousif b,
M.E. Elzain b, M. Shongwe c
a
Department of Mathematical and Physical Sciences, College of Arts & Sciences, University of Nizwa, P.O. Box 33, Code 616, Oman
b
Department of Physics, College of Science, Sultan Qaboos University, P.O. Box 36, Code 123, Al-khoud, Oman
c
Department of Chemistry, College of Science, Sultan Qaboos University, P.O. Box 36, Code 123, Al-khoud, Oman

a r t i c l e i n f o abstract

Article history: Compositions of polycrystalline Mg–Zn mixed ferrites with the general formula Mg1  xZnxFe2O4
Received 23 August 2011 (0 r x r 1) were prepared by the standard double sintering ceramic method. The structural properties
Received in revised form of these ferrites have been investigated using X-ray diffraction and infrared absorption spectroscopy.
10 December 2011
The lattice parameter, particle size, bonds length, force constants, density, porosity, shrinkage and
Accepted 13 December 2011
Available online 21 December 2011
cation distribution of these samples have been estimated and compared with those predicted
theoretically. Most of these values were found to increase with increasing Zn content. The energy
Keywords: dispersive (EDS) analysis confirmed the proposed sample composition. The scanning electron micro-
Mg–Zn Ferrites scope (SEM) and transmission electron microscope (TEM) micrographs showed aggregates of stacked
Spinel
crystallites of about 200–800 nm in diameter. Far infrared absorption spectra showed two significant
IR
absorption bands. The wave number of the first band, n1, decreases with increasing Zn content, while
XRD
Electrical resistivity the band, n2 shifts linearly towards higher wave numbers with Zn contents, over the whole composition
Force constants range. The room temperature electrical resistivity was found to decrease as Zn-content increases.
Vacancy model parameters Values of the vacancy model parameters showed that the packing factors Pa and Pb decrease, the
fulfillment coefficient, a, remains almost constant and the vacancy parameter, b, strongly increases
with increasing Zn content in the sample. The small values of Pa, Pb, a and the strong increase of the
vacancy parameter, b, indicate the presence of cation or anion vacancies and the partial participation of
the Zn2 þ vacancies in the improvement of the electrical conductivity in the Mg–Zn ferrites.
& 2011 Elsevier B.V. All rights reserved.

1. Introduction structural properties of the ferrite materials. They give informa-


tion about the positions of the ions in the spinel lattice, the
Ferrites have been the subject of extensive investigation vibration modes, the valance state of the ions and their occupa-
because of their wide range applications such as information tion in the spinel lattice crystal. Moreover, the IR spectra scans
storage systems, gas sensors, microwave devices and magnetic can be used to detect the completion of the solid state reaction,
recording and electronic industries. These materials are charac- cations distribution and the deformation of spinel structure. The
terized by high electrical resistivity and unique magnetic and IR spectra absorption bands mainly appear due to vibrations of
electrical properties, which result in low eddy currents and the oxygen ions with the cations producing various frequencies in
dielectric losses [1–3]. The crystalline structure of the spinel the unit cell. The vibration frequency depends on the cation’s
ferrites has two sites, the tetragonal A-sites and octahedral mass, cation–oxygen distance and the bonding force [5].
B-sites. These materials have the general chemical formula The vacancy model [6] describes the influence of vacancies and
(TM21 
þ 3þ 2þ 3þ
dFed ) [TMd Fe2  d]O
2
. The factor d represents the frac- mixed valence on the transport processes in solid solutions with

tion of Fe ions at the A-sites. Magnesium ferrite, MgFe2O4, is the spinel structure. This model deals with four quantities; the
predominantly an inverse spinel and the degree of inversion coefficients of ion packing at the tetrahedral and octahedral sites,
depends upon heat treatment [4]. The zinc ferrite, ZnFe2O4, has the fulfillment coefficient of the unit cell, which determines the
the normal spinel structure, in which the Zn2 þ ions occupy the degree of the ionic packing of the spinel structure, the difference
tetrahedral sites where the octahedral B-sites are occupied by the of Pauling electronegativities, which give an evaluation of the
Fe3 þ ions. The IR spectra are important tools to investigate the ionic bonds strength in the tetrahedral and octahedral sites and
the vacancy parameter, which is a measure of the total vacancy
concentration in the sample.
n
Corresponding author. Tel.: þ968 25446434; fax: þ968 225446289. Moreover, distinct relations were reported [7] to exist between
E-mail address: Kadhimm@unizwa.edu.om (K.A. Mohammed). the ionic packing factors at the tetrahedral and octahedral sites

0921-4526/$ - see front matter & 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.physb.2011.12.097
796 K.A. Mohammed et al. / Physica B 407 (2012) 795–804

and the electrical conductivity and the magnetic ordering at these


sites, respectively.
The aims of this study are to investigate the effect of Zn
substitution on X-ray parameters, IR absorption bands, room
temperature electrical resistivity and to test the validity of the
vacancy model on the relevant parameters of the mixed
Mg1  xZnxFe2O4 ferrites.

2. Experimental

Samples of the spinel series Mg1  xZnxFe2O4 (x ¼0–1 step 0.1)


were prepared using the conventional double sintering technique,
in which powders of high purity MgO, ZnO and Fe2O3 oxides were
mixed in required proportions in agate mortar. The mixture was
first sintered at 1000 1C for 24 h in air medium followed by
cooling to room temperature. The powder was then remixed and
ground once more to promote homogeneity. Pellets of 13 mm and
2–3 mm thick are prepared using hydraulic press of 10 tons/cm2.
These pellets and the rest of the powder were sintered at 1050 1C
for 24 h followed by natural cooling to room temperature. The
final total losses in sample masses were negligible (less than
0.1%). Moreover, the samples compositions were confirmed by
Mossbauer and the transmission electron microscope analysis.
The single-phase spinel structure was confirmed by the X-ray
diffraction spectrum of these samples. The X-ray diffraction
spectrum of these samples were examined using Phillips
PW1820 diffractometer with CuKa radiation (l ¼1.5404 Å). The
scan’s range was kept the same for all samples 2y ¼10–1001 using
a step size of 0.021 with sample time of 2 s.
Samples for recording IR spectra were prepared by mixing
small quantity of the powder of the samples with solid KBr. The
mixed powder of samples was then placed in a cylindrical disc and
pressed at 10 tons/cm2 by a hydraulic press. The IR measurements
of the prepared samples were recorded at room temperature in Fig. 1. The X-ray diffraction pattern for the Mg1  xZnxFe2O4 samples.
the range from 400 up to 1000 cm  1 using PERKIN-ELMER-1430
infrared Spectrophotometer.
Table 1
Samples for the DC resistance measurements at room tem- Values of rA, rB, aexp, ath and cation distribution of Mg1  xZnxFe2O4 ferrites.
perature were prepared by first grinding the surfaces to remove
any oxide, such as Fe2 þ , which might be caused by the distillation x aexp (Å) ra (Å) rb (Å) ath (Å) Cation distributions (Ref. [15])
off the sample surface, such as the Zn2 þ , during sintering at high
temperatures and to remove any contamination to the pellet 0.0 8.378 0.6697 0.6652 8.483 (Mg.03Fe.97)A [Mg.97Fe1.03]B
0.1 8.393 0.6840 0.6660 8.508 (Zn.1Mg.1Fe.8)A [Mg.8Fe1.2]B
surfaces during the pressing process [8]. The pellets surfaces were
0.2 8.405 0.6992 0.6664 8.532 (Zn.2Mg.08Fe.72)A [Mg.72Fe1.28]B
coated with a silver paste conductor to ensure a good electrical 0.3 8.409 0.7142 0.6669 8.556 (Zn.3Mg.08Fe.62)A [Mg.62Fe1.38]B
contact between the sample and the two copper terminals for the 0.4 8.417 0.7293 0.6674 8.581 (Zn.4Mg.07Fe.53)A [Mg.53Fe1.47]B
electrical resistance measurements. 0.5 8.421 0.7445 0.6678 8.605 (Zn.5Mg.05Fe.45)A [Mg.45Fe1.55]B
0.6 8.430 0.7595 0.6683 8.630 (Zn.6Mg.05Fe.35)A [Mg.35Fe1.65]B
The microstructure and sample morphology were examined
0.7 8.439 0.7746 0.6687 8.654 (Zn.7Mg.04Fe.26)A [Mg.26Fe1.74]B
with analytical scanning electron microscope (ASEM) model 0.8 8.442 0.7897 0.6692 8.679 (Zn.8Mg.035Fe.165)A [Mg.165Fe1.835]B
JSM-6510LA-JEOL and transmission electron microscope model 0.9 8.448 0.8047 0.6697 8.703 (Zn.9Mg.03Fe.07)A [Mg.07Fe1.93]B
JEM-1400-JEOL. 1.0 8.450 0.8200 0.6700 8.728 (Zn1)A [Fe2]B

3. Results and discussions parameters is attributed to the replacement of smaller Fe3 þ


(0.67 Å) ions with a larger ionic radius of the Zn2 þ (0.82 Å), at
The X-ray diffraction patterns for the Mg1  xZnxFe2O4 samples the tetrahedral sites, and the replacement of Mg2 þ ions (0.66 Å)
are shown in Fig. 1. The analysis of XRD patterns confirms the with a larger ionic radius Fe3 þ ion at the octahedral sites. The
formation of spinel cubic structure. The reflections from the nonlinear behavior of lattice parameter ‘‘aexp’’ with Zn content (i.e.
plains (1 1 1), (2 2 0), (3 1 1), (2 2 2), (4 4 0), (4 2 2), (5 1 1), violating Vegard’s law) was reported for other ferrite systems,
(6 2 0), (5 3 3), (6 2 2), (4 4 4) and (5 3 3) are shown by all which deviate from the complete normal or complete inverse
samples. This is an indication of the spinel cubic single-phase spinel type structure [9]. Values of the lattice parameter ‘‘aexp’’ for
formation. The lattice parameters ‘‘aexp’’ for all samples have been MgFe2O4 and ZnFe2O4 samples have been calculated as 8.378 and
determined from different diffraction lines of the XRD patterns 8.450 Å, respectively, which agree quite well with literature
employing the Winfit-XRD software program. Values of ‘‘aexp’’ are values of 8.387 [10], 8.384 [11] and 8.377 Å [12] for MgFe2O4
tabulated in Table 1(Fig. 3). as a function of the lattice parameter and 8.437 [13] and 8.33 Å [14] for ZnFe2O4. The larger value of the
with Zn content. It is clear that the lattice constant increases lattice parameter for ZnFe2O4 compared with that of MgFe2O4 is
smoothly with increase in Zn content. The increase in lattice due to the larger ionic radius of Zn2 þ than both Fe3 þ and Mg2 þ .
K.A. Mohammed et al. / Physica B 407 (2012) 795–804 797

Values of the experimental, ‘‘aexp’’, and theoretical, ‘‘ath’’, lattice 8.75 8.50
constants are shown in Table 1 (Fig. 3). It is clear that the ‘‘ath’’
8.70
values are greater than the ‘‘aexp’’ values. This is because the ‘‘ath’’
calculations assume a perfectly filled close packed spinel struc- 8.65
aexp

Lattice constants (A)


ture and the anions and cations are rigid spheres [6]. Similar 8.60 ath
behaviors have been shown by the Co–Cd Ferrites [15]. In the 8.45
Mg1  xZnxFe2O4 ferrites the Zn2 þ ions exclusively occupy the
8.55
tetrahedral sites, while most of the Mg2 þ ions prefer to occupy 8.50
the octahedral sites. The Fe3 þ ions are distributed among the
8.45
tetrahedral and octahedral sites. Thus the cation distribution for 8.40
the Mg1  xZnxFe2O4 ferrites can be written as 8.40

ðZn2x þ Mg2y þ Fe31xy


þ
ÞA ½Mg21xy
þ
Fe31 þ 8.35
þ x þ y B
8.30
The average cation radius at the tetrahedral and octahedral
sites, rA and rB, can be estimated for all samples employing the 8.25 8.35
cation distributions and the relations [16]: 0.0 0.2 0.4 0.6 0.8 1.0
Zn (x)
r A ¼ ðC AMg2 þ Þðr Mg2 þ Þ þ ðC AZn2 þ Þðr Zn2 þ Þ þ ðC AFe3 þ Þðr Fe3 þ Þ ð1Þ
Fig. 3. Variation of the lattice parameter, ‘‘aexp’’ and ‘‘ath’’ with x. (The inset: shows
ðC BMg2 þ Þðr Mg2 þ Þ þðC BZn2 þ Þðr Zn2 þ Þ þ ðC BFe3 þ Þðr Fe3 þ Þ variation of ‘‘aexp’’ with rA.)
rB ¼ ð2Þ
2
A B
where C and C are the ionic concentration at the tetragonal and
Table 2
octahedral sites, respectively, while r Mg2 þ , r Zn2 þ and r Fe2 þ are the
Values of u, d, aexp, Ra and Rb as a function x.
ionic radii of Mg2 þ , Zn2 þ and Fe3 þ ions, respectively. The cation
distributions have been obtained from the Mossbauer study [17]. x u d aexp (Å) RA (Å) ( 70.001) RB (Å) (7 0.001)
Table 1 shows results for the cation distributions, aexp, ath, rA and
rB calculations. Fig. 2 shows clearly that values of both rA and rB 0.0 0.3810 0.0060 8.378 1.901 2.045
increase with increasing the Zn2 þ ions content; this can be 0.1 0.3814 0.0064 8.393 1.910 2.046
0.2 0.3818 0.0068 8.405 1.919 2.046
assigned to the replacement of Fe3 þ ions with a larger ionic 0.3 0.3822 0.0072 8.409 1.925 2.043
radius of the Zn2 þ , at the tetrahedral sites, and the replacement of 0.4 0.3826 0.0076 8.417 1.933 2.042
Mg2 þ ions with a larger ionic radius Fe3 þ ions at the octahedral 0.5 0.3830 0.0080 8.421 1.940 2.040
sites. Similar results have been reported for the Co substituted 0.6 0.3834 0.0084 8.430 1.948 2.039
0.7 0.3838 0.0088 8.439 1.956 2.038
Ni–Zn ferrites [18].
0.8 0.3842 0.0092 8.442 1.962 2.036
The lattice parameter can be calculated using the following 0.9 0.3846 0.0096 8.448 1.970 2.034
relation, which relates the cation radius of the constituent 1.0 0.3850 0.0100 8.450 1.976 2.032
elements at the different lattice sites to the lattice parameter
[19]:
pffiffiffi
r A þRo þ 3ðr B þ Ro Þ 2.050
ath ¼ 8 pffiffiffi ð3Þ
3 3 1.98
where rA and rB are the cations radii of element at the tetrahedral
and octahedral sites, respectively and Ro is the oxygen ion radius 2.045
1.96
Bond Length (A)

( ¼0.138 Å) [20]. Table 1 and Fig. 3 show the calculated values of
the lattice parameter, ‘‘aexp’’ and ‘‘ath’’. It is clear that ‘‘ath’’ values
are higher than ‘‘aexp’’ values. The ‘‘ath’’ values vary linearly with 1.94 2.040

RB
1.92 RA
0.64 0.672 2.035

0.62
1.90
0.670 2.030
0.60
Cations Radii (A)

0.0 0.2 0.4 0.6 0.8 1.0


0.58 Zn (x)
rA
0.668
rB Fig. 4. Variation of the bond length, RA and RB, with x.
0.56

0.54
0.666 the increased Zn concentration in the sample. The inset in Fig. 3
0.52 shows the variation of the lattice constant ‘‘aexp’’ as a function of
the mean ionic radius, rA, at the tetrahedral site. It can be seen
0.50 0.664 that aexp values increase linearly with rA, verifying the suggested
0.0 0.2 0.4 0.6 0.8 1.0 linear relation between the two parameters [16]. These results
Zn (x) agree with results of the Ni–Zn and Ni–Cd ferrites [16].
The inter-ionic cation–anion distances (the bond lengths) at
Fig. 2. Variation of the ionic radii, rA and rB, with x. the A-sites, RA, and B-sites, RB, can be evaluated using the
798 K.A. Mohammed et al. / Physica B 407 (2012) 795–804

Table 3 3.3
Variation of Rx, Rx0 , Rx00 , dexp, dxrd, DWH and DXRD with x.

x Rx Rx 0 Rx00 dexp dXRD DWH DXRD 3.2 Rx


(Å) (Å) (Å) (g/cm3) (g/cm3) (nm) (nm)
3.1
0.0 3.104 2.820 2.964 2.80 4.16 33 58

Edge Length (A)


0.1 3.119 2.815 2.969 2.85 4.21 43 76
3.0 R x''
0.2 3.133 2.810 2.974 2.83 4.26 57 101
0.3 3.144 2.802 2.975 2.91 4.31 32 56
0.4 3.157 2.795 2.979 2.96 4.36 48 83
0.5 3.168 2.787 2.980 2.97 4.40 70 123 2.9
0.6 3.181 2.780 2.984 3.03 4.45 52 92 R x'
0.7 3.194 2.774 2.987 3.14 4.49 46 81 2.8
0.8 3.204 2.765 2.989 3.10 4.53 58 102
0.9 3.216 2.757 2.991 3.26 4.57 53 92
1.0 3.227 2.749 2.992 3.45 4.61 33 57 2.7

2.6
relations [21]
0.0 0.2 0.4 0.6 0.8 1.0
pffiffiffi
RA ¼ a 3ðd þ 18Þ ð4Þ Zn (x)
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Fig. 5. Variation of the bond lengths Rx, Rx0 and Rx00 with x.
2 d 1
RB ¼ a 3d  þ ð5Þ
2 16
Values of Rx, Rx0 and Rx00 are shown in Table 3 and Fig. 5. It is clear
d ¼ u0:375 ð6Þ
that values of Rx, Rx0 and Rx00 increase, decrease and increase
where d represents the deviation from the oxygen parameter uideal slowly, respectively, with Zn content. Similar results have been
( ¼0.375). Values of u for Mg1  xZnxFe2O4 samples were found by reported for the Cd–Zn [24] and Cd–Ni–Zn [25] ferrite systems.
interpolation between those reported earlier for pure magnesium The changes in u cause a distortion of the octahedral sites
and zinc ferrites, which are equal to 0.381 and 0.385 Å, respec- symmetry. This distortion results in shortening the octahedral
tively [21]. Values of RA (RB) increase (decrease) with increasing shared edge bringing the anions into close contact along those
zinc content (x). The changes in RA and RB values with x might be edges. The calculated results for the octahedral shared edges, Rx0 ,
caused by the substitution process. Similar results have been and the bond length, RB, for these samples support such explana-
reported for other ferrite systems, such as the Mn–Al ferrites [22]. tions. The average value of Rx0 ( ¼2.79 Å) corresponds to about
The increase in Zn þ 2 content leads to the A-site expansion and to twice that of the O2  ions ( ¼2.70 Å).
a relative displacement of the oxygen anions cause shrinkage of The interionic distances between the cation–cation (Me–Me)
the B-sites. Values of d, RA and RB are shown in Table 2 and Fig. 4. (b, c, d, e and f) and cation–anion (Me–O) (p, q, r and s) and the
It is clear that the bond lengths at tetrahedral sites, RA, are smaller angles between these distances in spinels are shown in Fig. 6 [26].
than those of octahedral sites, RB. This can be interpreted as more These distances are calculated for the Mg1  xZnxFe2O4 ferrites
covalent bonding of Fe3 þ ions at the B-sites than A-sites. These using the experimental values of lattice constant, aexp, and oxygen
results support the interpretation that correlates the decrease in positional parameter u using the following relations [27]:
the bond length to the increased covalent bonding characteristics
Cation-anion distances Cation-cation distances
[23]. The deviation of u from the standard value can be taken to
some extent as a measure of the triagonal distortion of the oxygen   apffiffiffi
p ¼ a 58 u b¼ 2
coordinates at the octahedral sites. This increased deviation   pffiffiffi a4pffiffiffiffiffiffi
reflects the increasing effect of the triagonal distortion at q ¼ a u 14 3 c ¼ 8 11
  pffiffiffiffiffiffi  pffiffiffi
the B-sites, which increases as x increased. Similar effect has r ¼ a u 14 11 d ¼ 4a 3
 pffiffiffi 3apffiffiffi
been reported for other systems such as Co substituted Ni–Zn s ¼ a 13 u þ 18 3 e¼ 8 3
ferrites [18].  pffiffiffi
f ¼ 4a 6
The increased values of d with Zn content (x) implies the
existence of a progressive expansion of the tetrahedral interstices The calculated values are shown in Tables 4 and 5. It can be
to accommodate the Zn2 þ ions which replace the Fe3 þ ions at the seen that all these distances (except p) increase slowly with
A-sites. The anions move away from the nearest tetrahedral increasing Zn content (x). The increases in the interionic distances
cations without changing the structure symmetry. The increasing are in accordance with increase in the unit cell volume. The bond
values of the mean ionic tetrahedral radii rA, the bond length RA angles y1, y2, y3, y4 and y5 are calculated applying simple
and the tetrahedral edge Rx are all a consequence of this expan- trigonometry principles using the values of the interionic dis-
sion [24]. tances [27]. The increase (decrease) of the angle values suggests
The tetrahedral edge length Rx, the shared octahedral length Rx0 strengthening (weakening) between the concerned cation–cation
and the unshared octahedral edge length Rx00 of these cubic mixed or cation–anion interactions. The distances effects are opposite to
spinel oxides have been calculated using the following equations that of the angles. It is seen that angles y1, y2 and y5 decrease
[24]: while angles y3 and y4 increase with Zn content (x). Accordingly
pffiffiffi the interactions between cation–anion (A–B) strengthen, while
Rx ¼ a 2ð2u0:5Þ ð7Þ the interactions between cations (B–B) weaken. But the increases
pffiffiffi in the distances values suggest weakening of these interactions.
Rx0 ¼ a 2ð12uÞ ð8Þ The overall resultant strength of the different magnetic interac-
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi tions decreases as the Zn content increases. These results are
11 consistent with the magnetic behaviors of these Mg–Zn ferrites
Rx00 ¼ a 4u2 3u þ ð9Þ
16 [11,17].
K.A. Mohammed et al. / Physica B 407 (2012) 795–804 799

Fig. 6. The interionic distances and angles between cations–cations and anions–cations at the tetrahedral and octahedral sites in spinel ferrites.

Table 4
Values of the interionic distances in (Å) as a function of Zn(x).

x p q r s b c d e f

0.0 2.0442 1.9010 3.6400 3.6568 2.9621 3.4733 3.6278 5.4417 5.1305
0.1 2.0445 1.9102 3.6577 3.6653 2.9674 3.4796 3.6343 5.4514 5.1396
0.2 2.0441 1.9187 3.6741 3.6725 2.9716 3.4845 3.6395 5.4592 5.1470
0.3 2.0417 1.9255 3.6870 3.6762 2.9730 3.4862 3.6412 5.4618 5.1494
0.4 2.0403 1.9331 3.7017 3.6816 2.9759 3.4895 3.6447 5.4670 5.1543
0.5 2.0379 1.9399 3.7146 3.6853 2.9773 3.4912 3.6464 5.4696 5.1568
0.6 2.0367 1.9478 3.7298 3.6912 2.9805 3.4949 3.6503 5.4754 5.1623
0.7 2.0355 1.9557 3.7449 3.6971 2.9836 3.4986 3.6542 5.4813 5.1678
0.8 2.0328 1.9623 3.7575 3.7003 2.9847 3.4999 3.6555 5.4832 5.1696
0.9 2.0309 1.9695 3.7713 3.7049 2.9868 3.5024 3.6581 5.4871 5.1733
1.0 2.0280 1.9758 3.7834 3.7077 2.9875 3.5032 3.6590 5.4884 5.1745

Table 5 4.7 3.6


Values of the interionic angles (in degrees) as a function of Zn(x).
3.5
4.6
x h1 h2 h3 h4 h5
3.4
0.0 123.34 144.95 92.85 125.92 74.48 4.5
Density (g/cm3)

0.1 123.22 144.38 93.05 125.96 74.14 3.3


0.2 123.08 143.83 93.25 126.01 73.80 4.4
0.3 122.96 143.29 93.45 126.05 73.46 3.2
d xrd
0.4 122.83 142.75 93.65 126.09 73.12
0.5 122.71 142.22 93.85 126.14 72.79 4.3 3.1
0.6 122.58 141.69 94.06 126.18 72.45
0.7 122.45 141.17 94.26 126.23 72.12 d exp
4.2 3.0
0.8 122.33 140.66 94.47 126.27 71.80
0.9 122.20 140.15 94.67 126.31 71.47 2.9
1.0 122.08 139.65 94.88 126.36 71.15 4.1
2.8
4.0
The X-ray density dXRD for each sample is calculated according 0.0 0.2 0.4 0.6 0.8 1.0
to the relation [28]: Zn (x)

ZM Fig. 7. Variation of the dexp and dXRD densities with x.


dXRD ¼ ð10Þ
Na3
where Z, M, N and a3 represent number of molecules per unit cell samples. This might be caused by the existence of pores in these
( ¼8), molecular weight, Avogadro’s number and volume of the samples. The nature of variation of dexp supports the variation in
unit cell, respectively. The bulk dexp densities of the specimens the lattice parameter. The difference between values of dexp and
were evaluated using the measured mass, m, and volume, V, of the dXRD data may be attributed to the porosity of the prepared
samples using the following relation: samples.
The porosity, p, was calculated according to the following
m
dexp ¼ ð11Þ equation:
V
dexp
Values of the calculated dXRD and dexp are shown in Table 3 and p ¼ 1 ð12Þ
dXRD
plotted in Fig. 7 as a function of Zn content in the sample. Both
dexp and dXRD densities increase with increasing Zn content. The The effect of Zn substitution on the porosity is shown in Fig. 8.
true density, dXRD, is higher than the bulk density, dexp, for all Zinc substitution increases the porosity (up to x¼0.6) thus
800 K.A. Mohammed et al. / Physica B 407 (2012) 795–804

0.34

0.32
Porosity

0.30

0.28

0.26

Fig. 10. EDS pattern of the surface of a solid piece of x¼ 0.5.


0.24
0.0 0.2 0.4 0.6 0.8 1.0
Zn (x)

Fig. 8. Variation of the porosity, p, with x (the inset shows 2r vs. x).

0.010

0.008
B Cos (0)

0.006

0.004
Fig. 11. SEM micrograph for the surface of a solid piece of x ¼0.5.

0.303 0.304 0.305 0.306 0.307 0.308


Sin (0)

Fig. 9. Variation of b cos(y) vs. sin(y) for Mg1  xZnxFe2O4 ferrites. (the inset:
b cos(y) vs. sin(y) for one sample).

increase the sample density. These data reveal clearly that with
increasing Zn concentration in the ferrite apparent porosity of the
samples increase because of increase in lattice parameter. Mean-
while the sample volume shrinkage increases as Zn content increase.
The inset in Fig. 8 shows the sample diameter as a function of Zn
content. The increased sample porosity and volume shrinkage with
increasing Zn content might be attributed to the difference in specific
gravity and hence the lattice parameter of the ferrite components.
Similar results have been reported for the Cd–Ni–Zn ferrites [25].
The average crystallite size, DXRD, of all samples were eval-
uated from the reflected X-ray diffraction peaks using Scherrer’s
equation [28]:
kl Fig. 12. TEM micrograph for crashed small solid piece of x ¼0.5.
DXRD ¼ ð13Þ
b cos y
where the constant k¼0.89, l is the wavelength of the X-ray Cd–Ni–Zn [25] ferrites. Moreover, the average particle sizes have
radiation ( ¼1.5404 Å), y is the diffraction angle and b is the full been estimated for each sample using the Williamson–Hall plot
width at half maximum (FWHM) of the most intense peak (3 1 1) [32] in which values of b cos(y) vs. sin(y) have been plotted. The
in radian. Values of the crystallite size are given in Table 5. The intercept of the fitted straight line equals kl/DWH:
average value for DXRD was observed to be about 48 nm. The
kl
average crystallite size might be higher than this because the b cos y ¼ þ2e sin y ð14Þ
DWH
width, b, is not corrected for the instrumental broadening.
Comparable results have been reported for the MgFe2O4 with where e is the strain introduced inside the sample. A typical
high sintering temperature [29], Ni–Cu–Zn [30], Co–Mg [31] and Williamson–Hall plot for one of the samples is shown in the inset
K.A. Mohammed et al. / Physica B 407 (2012) 795–804 801

Table 6 of the surface of a solid pellet is shown in Fig. 11. Aggregates of


Values of n1, n2, Kt, Ko and r as a function of Zn(x). nearly rounded to cubic stacked crystallites of about 200–800 nm
can be identified. Fig. 12 shows the TEM micrographs of a crashed
x n1 (cm  1) n2 (cm  1) Kt  105 Ko  105 q  108 (X-cm)
(dynes/cm) (dynes/cm) 7 105 small solid piece confirm the SEM results. The crystallites sizes
obtained from SEM and TEM micrographs are larger than those
0.0 571 432 2.372 1.358 34.5 obtained from the XRD scans using Scherrer’s equation and
0.1 572 432 2.381 1.358 13.6 Williamson–Hall plot. These differences might be caused by
0.2 571 434 2.372 1.370 4.30
0.3 570 436 2.364 1.383 1.91
ignoring the crystal strain or defects contributions to the line
0.4 568 438 2.347 1.396 3.00 broadening of the FWHM values. Grain sizes as large as 0.335,
0.5 566 436 2.331 1.383 2.85 1.44 and 41.9 mm for Mg–Zn ferrites were reported [33].
0.6 562 439 2.298 1.402 1.10 The infrared spectra measurements of the Mg1  xZnxFe2O4 in
0.7 562 438 2.298 1.396 0.67
the range 400–1000 cm  1 indicate the presence of two strong
0.8 562 440 2.298 1.409 0.35
0.9 564 446 2.314 1.447 0.28 absorption bands n1 (571–564 cm  1) and n2 (432–447 cm  1). The
1.0 564 447 2.314 1.454 1.97 band positions n1 and n2 are listed in Table 6 as a function of Zn
content. These bands are common features of all the spinel
ferrites [5]. The absorption band n1 corresponds to stretching
vibration mode of Fe3 þ –O2  in the tetrahedral sites, and n2
corresponds to the metal–oxygen vibration in octahedral sites,
Fe3 þ –O2  . Fig. 13 shows the typical IR spectra of the
Mg1  xZnxFe2O4 ferrites. The difference in the position of the
two strong bands can be related to the differences in the Fe3 þ –
O2  distances for A-sites and B-sites. It is clear (Table 2) that the
Fe–O bond lengths at the A-sites (1.901–1.976 A) is shorter than
that of the B-sites (2.045–2.032 A). This effect has been inter-
preted on the base of more covalent bonding of the Fe3 þ ions on
the A-sites than that on the B-sites [34]. The wave number of the
first band, n1, decreases with increasing Zn content. This variation
in the band position can be attributed to the increase in the
cation–oxygen (A–O) bond length. This indicates the weakening
of the metal–oxygen bonds at the tetrahedral sites due to the
replacement of the Fe3 þ ions with Zn2 þ ions at the tetrahedral
sites. The residence of Zn2 þ ions on the tetrahedral (A) site causes
an equivalent amount of Fe3 þ ions to migrate from A-site to
B-site, which results in an increase of the (A–O) bond length of
the tetrahedral site of spinel structure (i.e. decrease in n1 band).
On substitution of Zn2 þ ions the position of n2 band shifts almost
linearly towards higher wave numbers over the whole composi-
tion range. This indicates the strengthening of the metal–oxygen
bonds at the octahedral sites due to the shifts of more Fe3 þ ions
from the tetrahedral sites to the octahedral sites. These results
agree to some extent with results of the Cu–Zn mixed ferrites [35]
and the Cu1  xZnxFeCrO4 ferrites [36]. The presence of Fe2 þ ions
causes local deformation in the spinel lattices, which has been
attributed to the Jahn–Teller distortion effect. This effect causes
splitting or shoulders on the absorption bands. A third band, n3, of
lower intensity, closer to n2 (less than 400 cm  1) has been
reported by many authors, for the Zn rich compositions [37]. It
is clear that the wave numbers of n3 and n4 lie outside the range of
these measurements. A relatively weak shoulder appears in the
range 650 up to 700 cm  1. The intensity of this shoulder
Fig. 13. IR spectra for the Mg1  xZnxFe2O4 ferrites. increases with increasing Zn2 þ ions; this leads to narrowing of
the band n1. This shoulder might be caused by the presence of two
covalent states of the ions such as the Fe2 þ –O2  and Zn þ 2–O2 
of Fig. 9. Values of DWH are shown in Table 3. The average value bonds at the tetrahedral sites. Similar results have been reported
for DWH was found to be about 84 nm. for the alumina-doped zinc ferrites [38]. The broadening is
Values of b cos(y) vs. sin(y) for the main X-ray (3 1 1) plane commonly observed in the inverse spinel ferrites caused by the
have been plotted for all samples are shown in Fig. 9. The inset statistical distribution of Fe3 þ ions on A- and B-sites. Therefore,
shows b cos(y) vs. sin(y) for one sample. The general trend of data the bands are sharpened as one goes from the inverse spinel
variation in this figure supports the increase of the crystallite size sample, MgFe2O4, to the normal spinel sample, ZnFe2O4. Similar
as Zn content increase in the sample. results have been reported for the mixed spinel Ni1  xZnxFe2O4
The results of the analytical scanning electron microscope ferrites [39,40], which agree quite well with these results. It is
(ASEM) confirm the purity of samples, the lack of any impurities clear from the XRD calculations that the bond lengths of the ions
and the expected stoichiometry. It also shows a well-packed and present in the tetrahedral sites, Ra, increases while the bond
continuous grain structure with porosity and small holes at the lengths at the octahedral sites, Rb, decreases with increase in Zn
crystallites boundaries. A representative energy dispersive analy- contents, x. It is well known that the bond length is inversely
sis (EDS) graph is shown in Fig. 10 for x ¼0.5. The SEM micrograph proportional to the vibration frequency. This explains the increase
802 K.A. Mohammed et al. / Physica B 407 (2012) 795–804

Rb (A) stretching in the tetrahedral sites leads to higher force constant


2.02 2.03 2.04 than the octahedral site. Variations of Kt and Ko with RA and RB are
shown in Fig. 14. The increased (decreased) values of Ra (Rb) with
2.38 1.46 Zn content support the decrease (increase) in the force constants
at the A-sites (B-sites), respectively. Similar results have been
1.44 reported for the mixed Cu–Zn ferrites [35].

Ko (dyne/cm) x 105
Kt (dyne/cm) x 105

2.36 Values of the measured room temperature DC-electrical resis-


1.42 tivity for the Mg1  xZnxFe2O4 ferrites are shown in Table 4 and in
2.34 Fig. 15 as a function of Zn content. All samples showed high
1.40
resistivity of order of (108–109) O-cm at room temperature.
Similar results have been reported for the Mg-ferrites [42,31],
2.32 Mn-ferrites [43] and Ni–Zn ferrites [44]. It is clear that the
1.38
resistivity of these ferrites decreases with increase in Zn-content.
The addition of larger ionic sized Zn2 þ ions distorts the ferrite
2.30 1.36 lattice by firstly, increasing the lattice parameter, ‘‘a’’ and sec-
ondly, increasing the number of Fe2 þ and Zn þ ions and forcing x
1.90 1.95 2.00 2.05
Fe3 þ ions to migrate from A-site to B-site to replace x Mg2 þ ions.
Ra (A) These lead to the increase in the hopping processes for both
electrons and holes and to reduce the indirect interaction
Fig. 14. The force constants Kt and Ko vs. bond lengths, Ra and Rb.
between the Fe3 þ ions on the A- and B-sites. The motion of
charges will be affected by this distortion causing the resistivity
35 10.0 of these samples to decrease. There is an increase in the resistivity
for the sample Mg0.5Zn0.5Fe2O4 (x¼0.5). The structure of this
Log p (ohm-cm) x 10

9.5
30 sample is probably neither inverse nor normal. The sample
9.0 ZnFe2O4 (x ¼1) exhibits noticeable increase in its resistivity. The
p (ohm-cm) x 108

25 8.5 transition to the complete normal structure might be the reason


for this increase. The resistivity values for these ferrites are
20 8.0
considerably higher than those reported for Ni–Zn–Me (Me ¼Cu,
7.5 Cd, Co, Ca and Mn) ferrites [45], Co–Ni–Mn ferrites [40] and
15 Cd–Ni–Zn ferrites [25]. It is well known that Zn ions prefer the
7.0
0.0 0.2 0.4 0.6 0.8 1.0 occupation of tetrahedral A-sites while Mg and Fe ions partially
10
Zn (x) occupy A- and B-sites. As Zn content increases at the A-sites, the
Mg ions concentration at the B-sites will decrease. As a result
5
some of the Fe3 þ ions will be forced to migrate from A-sites to
0 B-sites. This leads to the increase in the electron hopping process
(Verwey mechanism) between Fe2 þ and Fe3 þ ions at the B-sites,
0.0 0.2 0.4 0.6 0.8 1.0 which causes the decrease in the electrical resistivity (increase in
Zn (x) conductivity) as the Zn content increases. The decrease in resis-
tivity value with increasing Zn content has been shown by other
Fig. 15. Room temperature electrical resistivity, r as a function of x. (The inset: ferrites such as Mn–Zn [43], Ni–Zn [8,44] and Zn–Mg [46] ferrites,
shows log r as a function of x). the presence of Zn ions enhanced the increase in the conduction
process. The inset in Fig. 15 shows values of log (r) vs. Zn content.
The two distinct regions might suggest that there are two well
(decrease) in n1 (n2) values. Similar results have been reported for defined system structures inverse and normal spinel separated by
the alumina doped zinc ferrite [38]. a narrow mixed region.
The force constants of the ions at the tetrahedral site (Kt) and The ionic packing coefficients Pa and Pb at the tetrahedral and
octahedral site (Ko) have been calculated by substituting the IR octahedral sites, respectively, can be estimated using the follow-
band frequencies n1 and n2 in the following standard formulae ing equations [7]:
[35,40]: pffiffiffi
r xt ¼ ðu0:25Þa 3Ro ð16Þ
K ¼ 4p2 c2 n2 m ð15Þ
where c is the light speed (  2.99  10 cm/s), n is the vibration 10 r xo ¼ ð0:625uÞaRo ð17Þ
frequency of the A- and B- sites and m is the reduced mass of the
Fe3 þ and O2  ions (  2.601  10  23 g). Table 6 shows variation of r xt
Pa ¼ ð18Þ
Kt and Ko with Zn content (x). It is clear that values of the force RA
constants Kt (Ko) decrease (increase) as Zn content (x) increases.
r xo
This behavior has been attributed to the variation in cation– Pb ¼ ð19Þ
RB
oxygen bond lengths at the A- (B-)sites. The decrease (increase) in
Kt (Ko) values can be related to the increase (decrease) in bond where rxt and rxo are the interstitial radii and RA and RB are the
length, Ra (Rb) at the tetrahedral (octahedral) sites. Longer bond average values of the ionic radii at the tetrahedral and octahedral
length requires less energy to break the bond. The increase in Ko sites, respectively, u is the anion parameter, ‘‘a’’ is the lattice
values as x increases can be attributed to the more transfer of parameter and Ro is the anion radius. It is claimed that [6] the
Fe3 þ ions from A-sites to the B-sites, which causes some kind of small values of the packing factors, Pa and Pb ( o1), testify to the
charge imbalance at the B-sites, which is likely to shift the oxygen smaller ion distances and larger overlapping of the cation and
ions towards the Fe3 þ ions causing the force constant, Ko (bond anion orbital, suggesting the existence of cation or anion vacan-
length, RB) to increase (decrease) as x increases [41]. The cies while the opposite holds for larger values of Pa and Pb ( ffi 1).
K.A. Mohammed et al. / Physica B 407 (2012) 795–804 803

Table 7 (decrease in r values) in these mixed ferrites. These results agree


Values of the ionic packing, fulfillment, vacancy and Pauling electronegativity with results of the thermoelectric power of the Mg–Zn ferrites
coefficients of Mg1  xZnxFe2O4 ferrites.
[46–49] in which the presence of both negative and positive
x Pa Pb a b (%) DvA DvB Seebeck coefficients were reported, suggesting that both n-type
rxt (Å) rxo (Å)
and p-type carriers are responsible for the conduction process in
0.0 0.5210 0.6642 0.778 0.998 0.649 3.668 2.986 3.177 these materials as follows:
0.1 0.5302 0.6645 0.775 0.998 0.647 4.000 3.000 3.170
0.2 0.5387 0.6641 0.770 0.997 0.646 4.399 3.002 3.166 Fe3 þ þ e 3Fe2 þ
0.3 0.5455 0.6617 0.764 0.992 0.646 5.066 3.006 3.162
0.4 0.5531 0.6603 0.758 0.989 0.646 5.625 3.010 3.158 Zn2 þ 3Zn þ þ e þ
0.5 0.5599 0.6579 0.752 0.985 0.646 6.279 3.012 3.155
0.6 0.5678 0.6567 0.748 0.983 0.646 6.793 3.016 3.150 Similar role has been claimed for Zn vacancies in the Mg–Zn
0.7 0.5757 0.6555 0.743 0.980 0.645 7.270 3.019 3.146 [50], ZnCr2  xNixSe4 [50] and ACr2  4 (A ¼Zn, Cd, Mn, X¼O, S, Se)
0.8 0.5823 0.6528 0.737 0.975 0.646 7.971 3.023 3.142
0.9 0.5895 0.6509 0.733 0.972 0.646 8.535 3.027 3.138
[6] spinels.
1.0 0.5958 0.6480 0.727 0.967 0.648 9.254 3.028 3.135 It has been reported [7] that a distinct relation exists between
the ionic packing factor of the metals at the tetrahedral site in
normal ferrites and their electrical conductivity type. The sam-
The fulfillment coefficient of the unit cell, a, determines the ples, Mg1  xZnxFe2O4 (except ZnFe2O4), are mixed spinel. The
degree of the ionic packing of the spinel structure. This coefficient packing factor, Pa and Pb, for the normal spinel sample; ZnFe2O4
can be calculated using the following equation [6]: was calculated and found to be equal 0.727 and 0.967, respec-
tively. The Pa value lies approximately at the border between the
32pðR3A þ 2R3B þ 4R3o Þ
a¼ ð20Þ insulator–semiconductor conducting regions in the proposed
3V
classification. The measured electrical resistivity of the ZnFe2O4
where V is the volume of the unit cell ( ¼a3exp ). sample (conductivity 5.08  10  7 O  1-cm  1) supports this
The vacancy parameter, b, is defined as a normalized volume classification.
of the missing ions at the nodal points of the spinel structure. It is
a measure of the total vacancy concentration existing in the
sample. It can be calculated using the following equation [6]: 4. Conclusions
a3th a3exp
b¼ n100% ð21Þ The analysis of the X-ray diffraction spectrum showed that the
a3th
Mg1  xZnxFe2O4 system, which has been prepared by the conven-
The difference in the Pauling electronegativities, Dw is an tional solid state reaction technique with double sintering at
evaluation of the ionic bonds strength in the tetrahedral and temperatures around 1000 1C is pure single-phase ferrite system.
octahedral sites. These differences in the electronegativity per Values of the theoretically and experimentally calculated lattice
cation at the A- and B-sites can be estimated (wmg ¼1.31, parameter, ‘a’, increase with Zn content in the sample. The
wZn ¼1.65, wFe ¼1.83 and wO ¼3.44) using the following equations variation was nonlinear for the experimental values of ‘a’. With
[6]: the exception of the volume shrinkage, some bonds length and
wA grain sizes the other lattice parameters increase with increasing
DwA ¼ wo  ð22Þ Zn concentration. The estimated values agree quite well with
4
those predicted theoretically. The EDS, SEM and TEM analysis of
wB the surface of a solid piece showed the existence of aggregates
DwB ¼ wo  ð23Þ
6 of stacked nearly rounded to cubic crystallites of about
where wA and wB represent the Pauling electronegativities of the (200–800) nm in diameter. The infrared spectra of these ferrites
cations at the A- and B-sites, respectively, and wo represent the give rise to two most prominent absorption envelopes. The high
oxygen electronegativity. The wA and wB values depend on the frequency band, n1, lies in the range between 564 and 571 cm  1
kind of cations and anions at the A- and B-sites [6]: was assigned to the Fe3 þ –O2  and Zn2 þ –O2  stretching vibra-
wA ¼ ðC AMg2 þ ÞðwMg2 þ Þ þ ðC AZn2 þ ÞðwZn2 þ Þ þðC AFe3 þ ÞðwFe3 þ Þ ð24Þ tions at the tetrahedral sites. The second main absorption band,
n2, is present in the range between 432 and 447 cm  1, which is
assigned to the Fe3 þ –O2  stretching vibrations at the octahedral
ðC BMg2 þ ÞðwMg þ 2 Þ þ ðC BZn2 þ ÞðwZn þ 2 Þ þðC BFe3 þ ÞðwFe3 þ Þ
wB ¼ ð25Þ sites. The absorption bands n1 and n2 revealed the formation of
2 single-phase spinel structure with two sublattices: the tetrahe-
where wmg ¼1.31, wZn ¼1.65, wFe ¼1.83 and wO ¼3.44. Calculations dral (A) and octahedral (B) sites. These results showed that the
of Pa, Pb, a, b, DwA and DwB coefficients for the Mg1  xZnxFe2O4 normal mode of vibration of tetrahedral clusters is higher (shorter
ferrites using Eqs. (16)–(25) are shown in Table 7. bond length) than that of octahedral clusters (longer bond
It is clear that values of the ion packing factor Pa decrease more length). Calculated values of the bond lengths RA and RB and ionic
strongly than Pb and the fulfillment coefficient a values remain radii rA and rB support this interpretation. The force constants of
almost constant with increasing Zn content. The small values the tetrahedral site, Kt, decrease and the octahedral site, Ko,
( o1) of Pa, Pb, a and the strong increase of the vacancy parameter, increase with increase in Zn content x. This behavior has been
b indicate the presence of cation or anion vacancies and the attributed to the variation in cation–oxygen bond length and the
domination of the Zn2 þ vacancies at the tetrahedral sites. The charge imbalance at the concerned sites. We can conclude that
electrical conductivity in ferrites was explained on the basis of the the gradual increase of the Zn content, x, in the studied
Verwey mechanism, i.e. exchange of electrons between the Mg1  xZnxFe2O4 system leads to gradual transformation from the
adjacent Fe2 þ and Fe3 þ ions that are distributed randomly over inverse to normal spinel structure. The system transfer from an
the octahedral sites. Since the hopping mechanism is the most almost inverse spinel MgFe2O4 (x¼0) to normal spinel ZnFe2O4
probable conduction mechanism in the Mg–Zn ferrites the pre- (x ¼1.0) leads to sharpening of the absorption bands. The room
sence of the Zn vacancies in the Mg–Zn ferrites might be partially temperature electrical resistivity is of order of (108–109)O-cm.
responsible for the improvement of the electrical conductivity The increase (decrease) of Zn (Mg) content lowers (raises) the
804 K.A. Mohammed et al. / Physica B 407 (2012) 795–804

resistivity in these ferrites. The ionic packing factors Pa and Pb [19] S.A. Mazen, M.H. Abdallah, B.A. Sabrah, H.A.M. Hasham, Phys. Status Solidi
decrease, the fulfillment coefficient, a, remains almost constant A 134 (1992) 263.
[20] R.D. Shannon, Acta Crystallogr. A 32 (1976) 751.
and the vacancy parameter, b, strongly increases as the Zn [21] K.J. Standley, Oxide Magnetic Materials, Claredon Press, Oxford, 1990.
content increase in the sample. The small values ( o1) of Pa, Pb, [22] R.L. Dhiman, S.P. Taneja, V.R. Reddy, Adv. Condens. Matter Phys. 2008 (2008)
a and the strong increase of the vacancy parameter, b indicate the 1. Article ID 703479.
[23] B.F. Levine, Phys. Rev. B 87 (1973) 2591.
presence of cation or anion vacancies and the domination of the
[24] C. Otero Arean, E. Garcia Diaz, J.M. Rubio Gonzalez, M.A. Villa Garcia, J. Solid
Zn2 þ vacancies at the tetrahedral sites. The presence of the Zn2 þ State Chem. 77 (1988) 275.
vacancies in the Mg–Zn ferrites might be partially responsible for [25] M. Siva Ram Prasad, B.B.V.S.V. Prasad, B. Rajesh, K.H. Rao, K.V. Ramesh,
the improvement of the electrical conductivity in these mixed J. Magn. Magn. Mater. 323 (2011) 2115.
[26] A. Goldman, Modern Ferrite Technology, 2nd ed., Springer Science þ Business
ferrites. Media, Inc., New York, 2006.
[27] V.K. Lakhani, T.K. Pathak, N.H. Vasoya, K.B. Modi, Solid State Sci. 13 (2011)
539. (and references there in).
Acknowledgment [28] B.D. Cullity, Elements of X-ray Diffraction, Addison Wesley, 1959.
[29] Y. Huang, Y. Tang, J. Wang, Q. Chen, Mater. Chem. Phys. 97 (2006) 394.
[30] W.-C. Hsu, S.C. Chen, P.C. Kuo, C.T. Lie, W.S. Tasi, J. Mater. Sci. Eng. B 111
One of us (Kadhim Ahmed) would like to thank Dr. C.B. Kolekar (2004) 142.
for useful comments. [31] M.A. Ahmed, A.A. EL-Khawlani, J. Magn. Magn. Mater. 321 (2009) 1959.
[32] G.K. Williamson, W.H. Hall, Acta Metall. 1 (1953) 22.
[33] S.S. Khot, N.S. Shinde, B.P. Ladgaonkar, B.B. Kale, S.C. Watawe, J. Adv. Appl. Sci.
References Res. 2 (2011) 460. (and references there in).
[34] B. Evans, S. Hanfner, J. Phys. Chem. Solids 29 (1968) 1573.
[35] H.M. Zaki, H.A. Dawoud, Physica B 405 (2010) 4476.
[1] E.C. Snelling (Ed.), Soft Ferrites, Properties and Applications, 2nd ed., Butter-
[36] M.C. Chhantbar, U.N. Trivedi, P.V. Tanna, H.J. Shah, R.P. Vara, H.H. Joshi,
worth Publishing, London, 1989. (and references therein).
[2] J.G. Lee, J.Y. Park, Y.-J. Oh, C.S. Kim, J. Appl. Phys. 84 (1988) 2801. K.B. Modi, Indian J. Phys. A 78 (2004) 321.
[3] B.P. Ladgonkar, P.N. Vasambekar, A.S. Vaingankar, J. Magn. Magn. Mater. [37] H.A. Dawood, S.K. Shaat, J. Al-Aqsa Uni. 10 (2006) 247.
210 (2000) 289. [38] N.M. Deraz, J. Anal. Appl. Pyrol. 91 (2011) 48.
[4] H. Knoch, H. Dannheim, Phys. Status. Solidi A 37 (1976) K135. [39] A.T. Raghavender, N. Biliskov, Z. Skoko, Mater. Lett. 65 (2011) 677.
[5] R.D. Waldron, Phys. Rev. 99 (1955) 1727. [40] P.A. Shaikh, R.C. Kambale, A.V. Rao, Y.D. Kolekar, J. Alloys Compd. 492 (2010)
[6] T. Gron, Philos. Mag. B 70 (1994) 121. 590.
[7] T. Satoh, T. Tsushima, K. Kudo, Mater. Res. Bull. 9 (1974) 1297. [41] K.B. Modi, U.N. Trivedi, P.U. Sharma, V.K. Lakhani, M.C. Chhantbar, H.H. Joshi,
[8] U. Ghazanfar, S.A. Siddiqi, G. Abbas, J. Mater. Sci. Eng. B 118 (2003) 132. Indian J. Pure. Appl. Phys. 44 (2006) 165.
[9] S.T. Alone, Sagar E. Shirsath, R.H. Kadam, K.M. Jadhav, J. Alloys Compds. [42] K S Rane, V.M.S. Verennkar, P.Y. Sawant, Bull. Mater. Sci. 24 (2001) 323.
509 (2011) 5055. [43] D. Ravinder, K. Latha, Matter. Lett. 41 (1999) 247.
[10] V.K. Mittal, P. Chandramohan, S. Bera, M.P. Srinivasan, S. Velmurugan, [44] A.M. Abdeen, J. Magn. Magn. Mater. 185 (1998) 199.
S.V. Narasimhan, Solid State Commun. 137 (2006) 6. [45] E. Rezlescu, L. Sachelarie, P.D. Popa, N. Rezlescu, IEEE Trans. Magn. 36 (2000)
[11] M.M. Haque, M. Huq, M.A. Hakim, Physica B 404 (2009) 3915. 3962.
[12] S.M. Kadam, S.I. Patil, S.H. Patil, B.K. Chougule, Bull. Mater. Sci. 15 (1992) 127. [46] H.M. Zaki, Physica B 404 (2009) 3356.
[13] C. Upadhyay, H.C. Verma, S. Anand, J. Appl. Phys. 95 (2004) 5746. [47] B.P. Ladgaonkar, P.N. Vasambekar, A.S. Vaingankar, Bull. Mater. Sci 23 (2000)
[14] M. Chakrabarti, D. Sanyal, A. Chakrabarti, J. Phys.:Condens. Matter 19 (2007) 1. 87.
[15] O.M. Hemeda, M.M. Barakat, J. Magn. Magn. Mater. 223 (2001) 127. [48] M.U. Islam, A.Y. Abbasi, T. Abbas, M.A. Chaudhry, A.Z. Chaudhry, J. Res. Sci.
[16] A. Globus, H. Pascard, V. Cagan, J. Physique Colloq. 38 (1977). C1-163-168. 14 (2003) 103.
[17] A.M. Gismelseed, K.A. Mohammed, H.M. Widatallah, A.D. Al-Rawas, [49] P.P. Hankare, V.T. Vader, U.B. Sankpal, L.V. Gavali, R. Sasikala, I.S. Mulla, Solid
M.E. Elzain, A.A. Yousif, J. Phys.: Conf. Ser. 217 (012138) (2010) 1. State Sci. 11 (2009) 2075.
[18] M.A. Amer, A. Tawfik, A.G. Mostafa, A.F. El-Shora, S.M. Zaki, J. Magn. Magn. [50] H. Duda, I. Jendrzejewska, T. Gron, S. Mazur, P. Zajdel, A. Kita, J. Phys. Chem.
Mater. 323 (2011) 1445. Solids 68 (2007) 80.

You might also like