Sei sulla pagina 1di 7

J Supercond Nov Magn DOI 10.

1007/s10948-013-2305-2

O R I G I N A L PA P E R

X-Ray Diffraction and Cation Distribution Studies in Zinc-Substituted Nickel Ferrite Nanoparticles
D.V. Kurmude R.S. Barkule A.V. Raut D.R. Shengule K.M. Jadhav

Received: 16 April 2013 / Accepted: 28 June 2013 Springer Science+Business Media New York 2013

Abstract Structural and cation distribution studies on Ni1x Znx Fe2 O4 (with x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0) ferrite nanoparticles by using X-ray diffraction analysis are reported. In this work the NickelZinc ferrites nanoparticles are synthesized by solgel auto combustion using respective metal nitrates and citric acid as fuel for the auto combustion reaction. Formation of ferrite nanoparticles having single-phase spinel structure is evident from the obtained X-ray diffraction patterns. Lattice constant values of the Ni1x Znx Fe2 O4 ferrite system are found to increase with increase of zinc substitution x . Broad and intense XRD peaks in the patterns indicate the nanocrystalline nature of the produced ferrite samples. Average particle size calculated from most intense Braggs reection (311) using Debye Scherrers formula is found to be 30 nm. The particle size is found to decrease with increase in zinc substitution x . Observed X-ray density is found to decrease with increase in zinc substitution x . Bulk density, porosity, and unit cell volume are also calculated from the XRD data. Distribution of metal cations in the spinel structure estimated from X-ray diffraction data show that along with Ni2+ ions most of the Zn2+ ions also occupy the octahedral [B] sites, which are attributed to nanosize dimensions of the ferrite samples.

Keywords NiZn nanoferrites Structural and physical properties Cation distribution

1 Introduction Cubic spinel ferrite is a group of technologically important materials having applications from microwave to radio frequencies. The spinel ferrite has general formula of MFe2 O4 , where M is any divalent ion of metals such as nickel, cadmium, zinc, magnesium, copper, etc. [1]. Structural, electrical, and magnetic properties of these materials effectively depend upon their stoichiometry, methods of synthesis, and cationic distributions among the available (A) and [B] sites of the face-centered cubic (fcc) spinel structure formed by oxygen anions at the corners. The unit cell of the spinel structure is obtained by doubling approximately face-centered cubic oxygen sublattice along each of the three directions. Of the resulting 64 tetrahedral (A) sites and 32 octahedral [B] sites, only 8 and 16 are occupied, respectively, by cations in stoichiometric spinel. The majority of spinel compounds belong to the space group 4 3 Fd3m (F1 /d 2/m , No. 227 in the International Tables) [2]. Occupation of the tetrahedral site entirely with a divalent transition metal produces a normal spinel structure, while occupation of the octahedral site with the divalent transition metal yields an inverse spinel structure. If divalent transition-metal ions are present on both (A) and [B] sublattices, the structure is of mixed or disordered spinel [3]. Occupation of metal ions at tetrahedral and octahedral sites also depends on method of preparation. For example, bulk nickel ferrite shows completely inverse spinel structure by occupying octahedral sites with nickel ions, whereas in nanocrystalline form small fraction of nickel ions is found to exist at available tetrahedral sites in a spinel structure.

D.V. Kurmude Milind College of Science, Aurangabad, 431004 M.S., India R.S. Barkule A.V. Raut K.M. Jadhav (B) Department of Physics, Dr. Babasaheb Ambedkar Marathwada University, Aurangabad, 431004 M.S., India e-mail: drjadhavkm@gmail.com D.R. Shengule Vivekanand Arts, Sardar Dalipsingh Commerce and Science College, Aurangabad, 431004 M.S., India

J Supercond Nov Magn

Knowledge of cation distribution and spin alignment is essential to understand the magnetic properties of spinel ferrites. The interesting physical and chemical properties of ferrospinels arise from their ability to distribute the cations among the available tetrahedral (A) and octahedral [B] sites. Determination of cation distribution at the tetrahedral (A) and octahedral [B] sites has been a subject of many studies. The cation distribution in spinel ferrites can be obtained from the analysis of various data obtained from Xray diffraction [4], Mossbauer spectroscopy [5], Magnetization measurements, [6], Electron spin resonance (ESR) [7], Neutron diffraction [8], Thermoelectric [9], and Nuclear magnetic resonance (NMR) [2]. Quantum mechanical method [10], Rietveld renement [11], and the reex program [12] can also be employed to determine the cation distribution in spinels. Methods suggested by Bertaut [13] and Furuhashi [14] are based on a comparison between the Xray diffraction intensities observed experimentally and those calculated for a large number of hypothetical crystal structures. Wet chemically synthesized polycrystalline spinel ferrites normally consist of ne particles, and they exhibit unusual physical properties, as compared to their bulk counterparts synthesized by conventional ceramic technique [15, 16]. Cation distribution is found to vary with method of synthesis, valence of cations doped, crystallite size, temperature, etc. Substituted nickel ferrites have been the subject of extensive investigation because of their microwave applications. Many reports are available on pure and substituted nickel ferrites synthesized by ceramic technique in bulk form and by wet chemical methods such as coprecipitation, hydrothermal, citrate precursor method, etc. in nanosize form [1719]. On going through these reports, it is seen that no systematic investigation is carried out on the zinc-substituted nickel ferrite nanoparticles synthesized by solgel auto combustion method. Effect of zinc and effect of nanosize form on the structural and physical properties along with cation distribution of nickel ferrite nanoparticles obtained by solgel auto combustion technique using citric acid as fuel is rarely studied. In this view, the structural, physical, and cation distribution studies of zinc-substituted nickel ferrites synthesized by solgel auto combustion method are undertaken, and the results obtained are reported in this paper.

from the solutions [21]. Aqueous solutions of nitrates were prepared in minimum amount of deionized water. The citric acid solution prepared separately as per the desired stoichiometry was then gradually added to the nitrate solutions, and the mixture was stirred continuously for half an hour to have sol. The pH of the sol was adjusted to 7 by dropwise addition of ammonia to it. In addition to continuous stirring, the sol is heated at 80 C in order to convert it into viscous brown gel. After the formation of gel the stirring is stopped, and gel is allowed to burn via auto combustion reaction forming the required loose oppy powder of the end product, the ferrite [22]. As prepared ferrite powder was then rst heated at 150 C for 2 h to remove water contents in it. After natural cooling, the powder is sintered at 600 C for 8 h in order to get the nanocrystalline ferrite powders. Flow chart of the synthesis procedure leading to formation of zinc-doped nickel ferrite is given in Fig 1. A typical chemical reaction (with x = 0.2) leading to the nal product of zinc substituted nickel ferrite can be stated as [23] 0.8Ni(NO3)2 + 0.2Zn(NO3)2 + 2Fe(NO3)3 + Ni0.8 Zn0.2 Fe2 O4 + 20 C6 H8 O7 9

40 80 CO2 + H2 O + 4N2 3 9

X-ray diffraction study of all NiZn ferrite samples is done using PANALYTICAL X-ray Diffractometer (Model: Xpert PRO MPD) in 2 range from 20 to 80 in step of 0.02 at room temperature. The cation distribution in the present ferrite system is determined from XRD data using Bertaut method.

3 Results and Discussion X-ray diffraction patterns corresponding to Ni1x Znx Fe2 O4 ferrite system for x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0, shown in Fig 2, conrm the formation of single-phase cubic spinel structure for all the samples. All the planes present in XRD pattern were indexed by combining Braggs law with plane spacing equation for cubic system [24]. Most of the planes such as (220), (311), (222), (400), (422), (511), and (440) belonging to the cubic spinel structure are found to be present in the XRD pattern of the samples under investigation. Moreover, a typical XRD pattern of pure nickel ferrite shows best match with PDF No. 10-325 that of the cubic spinel. The existence of broad peaks in the XRD patterns indicates the nanosize dimensions of the prepared ferrite particles [25]. The lattice constant was calculated using the relation a=d h2 + k 2 + l 2 , (1)

2 Experimental Nickelzinc ferrite samples with compositional formula Ni1x Znx Fe2 O4 (with x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0) were synthesized by solgel auto combustion method using AR grade nitrates of nickel, zinc, and ferric [20]. Citric acid is used as fuel. Using citric acid as fuel helps the homogeneous distribution of the metal ions to get segregated

J Supercond Nov Magn Fig. 1 Flow chart of the solgel auto combustion synthesis of Ni1x Znx Fe2 O4 ferrite nanoparticles

where a is the lattice constant, d is the interplanar spacing, and h, k , and l are the Miller indices. The interplanar spacing d is calculated by using the well-known Bragg law of X-ray diffraction, viz., n = 2d sin , (2)

where n is the order of diffraction, is the wavelength of the X-ray employed, and is Braggs glancing angle. The X-ray density dx is calculated using the relation dx = 8M , NA a 3 (3)

where M is the molecular weight of the composition, NA is the Avogadro number, and a is the lattice constant [26]. The percent porosity is determined using the equation P = 1 db , dx (4)

where db is the bulk density. The particle size t is calculated using the DebyeScherrer formula t=
Fig. 2 X-ray diffraction patterns of ferrite system Ni1x Znx Fe2 O4 with x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0

0.9 , cos

(5)

where is the full width at half maximum (FWHM) of the recorded XRD pattern, taken in radians. Most intense peak

J Supercond Nov Magn

(311) is considered for the calculation of particle size as it is the strongest peak appearing at lower diffraction angles and can be conveniently analyzed by computer t. Using the values of lattice parameter a , the unit cell volume is also determined for each sample. The obtained values of the structural parameters are tabulated in Table 1. 3.1 Effect of Zinc Substitution on Structural Parameters From Table 1 it is clear that the lattice constant a increases with zinc substitution as expected. The increasing behavior of lattice constant with Zn2+ content is because of replacement of Ni2+ ions having comparatively smaller ionic radius (0.69 ) by Zn2+ ions with larger ionic radius (0.74 ). The variation of the lattice parameter a as a function of Zn2+ ion concentration x in the Ni1x Znx Fe2 O4 matrix follows the Vegard law [27]. The values of obtained lattice constant are compared with that generated from the data of the bulk Ni Zn ferrite samples [28, 29]. These values are also presented in Table 1. The values are found to be lower than that of bulk lattice parameter except for x = 0.8. Variation of lattice constant a with Zn content x is shown in Fig 3. The X-ray density of pure nickel ferrite is found to be 5.391 g/cm3 , which matches closely with its reported value 5.38 g/cm3 [30]. The X-ray density of NiZn ferrites is found to decrease with zinc substitution from 5.391 g/cm3 to 5.320 g/cm3 . The X-ray density depends upon the lattice constant, which is increased with the increase in Zn concentration, so the corresponding X-ray density decreased with the increase in Zn concentration. A similar trend has been reported in studies of zinc-substituted manganese ferrite [31]. The values of the bulk density are found to be lower than those of the X-ray density values and are attributed to the presence of pores formed and developed at the time of sample preparation and sintering process [32]. The change in the density of the prepared ferrite samples is also attributed to the change in sintering conditions, sintering temperature, particle size, and chemical compositions. The X-ray density, bulk density, porosity, and specic surface area values are in agreement with the reported data [33, 34].
Table 1 Lattice constant (a ), X-ray density (dx ), bulk density (db ), porosity (P ), particle size (t ), and unit cell volume obtained for the Ni1x Znx Fe2 O4 spinel ferrite system

Different structural parameters such as hoping lengths (LA and LB ), tetrahedral bond length (dAX ), octahedral bond length (dBX ), tetrahedral edge length (dAXE ), shared octahedral edge length (dBXE ), unshared octahedral edge length (dBXEU ), tetrahedral site radius (rA ), and octahedral site radius (rB ) are determined by using the usual relations [35] and summarized in Table 2. The value of oxygen positional parameter is taken as 0.375 ; ionic radii of Ni2+ (0.69 ), Zn2+ (0.74 ), Fe3+ (0.645 ), and O2 (1.32 ) ions are considered as given by Shannon [36]. 3.2 Effect of Zinc Substitution on Particle Size Particle size of pure nickel ferrite is obtained to be 41 nm. Upon substitution of zinc, it was found to be reduced to average value of 30 nm. The presence of zinc obstructs the crystal growth in spinel ferrites. The crystal growth in the solution depends on various parameters, the most important one being the molecular concentration of the material approaching the surface of the tiny crystal during the growth process. Because of the liberation of latent heat at the surface, the local temperature is normally higher than the solution temperature. The surface temperature affects the molecular concentration at the surface of the crystal and, hence, the crys-

Fig. 3 Variation of lattice constant of Ni1x Znx Fe2 O4 spinel ferrite system with zinc content x in bulk and nanoform

aexp () Nano Bulk 8.340 8.365 8.388 8.408 8.426 8.442

dx (gm/cm3 ) 5.391 5.367 5.357 5.334 5.320 5.344

db (gm/cm3 ) 2.062 2.233 1.964 1.917 1.761 2.005

P (%) 61.75 58.40 63.35 64.07 66.89 62.48

t (nm) 41 32 27 30 30 34

V (3 ) 578 583 588 594 598 599

0.0 0.2 0.4 0.6 0.8 1.0

8.328 8.356 8.377 8.405 8.428 8.431

J Supercond Nov Magn Table 2 Hopping lengths (LA and LB ), tetrahedral bond length (dAX ), octahedral bond length (dBX ), tetrahedral edge length (dAXE ), shared octahedral edge length (dBXE ), unshared octahedral edge length X 0 0.2 0.4 0.6 0.8 1 LA 2.515 2.522 2.527 2.534 2.540 2.541 LB 2.436 2.443 2.448 2.455 2.461 2.461 dAX 1.890 1.896 1.901 1.907 1.912 1.913 dBX 2.033 2.040 2.045 2.052 2.058 2.058 dAXE 3.086 3.096 3.104 3.114 3.123 3.124 dBXE 2.803 2.812 2.820 2.829 2.837 2.838 (dBXEU ), tetrahedral site radius (rA ), octahedral site radius (rB ), and theoretical lattice constant (ath ) for the Ni1x Znx Fe2 O4 spinel ferrite system dBXEu 2.946 2.956 2.963 2.973 2.981 2.983 rA 0.570 0.576 0.581 0.587 0.592 0.593 rB 0.712 0.719 0.724 0.731 0.736 0.737 rA(cal.) 0.645 0.655 0.657 0.654 0.649 0.647 rB(cal.) 0.668 0.668 0.672 0.678 0.686 0.692 ath () 8.325 8.341 8.354 8.367 8.379 8.392

tal growth. The formation of zinc ferrite is more exothermic as compared to formation of nickel ferrite. Thus, it is expected that if one introduces zinc in the system, more heat will be liberated, decreasing the molecular concentration at the crystal surface and hence obstructing the grain growth [28] and reducing the particle size. Moreover, the reduction in particle size can be related to the electronic conguration of nickel (3d8 ) and zinc (3d10 ); obviously, as compared to zinc, nickel, which has incomplete electronic conguration, has more tendencies to interact with legends and O2 anions. The lack of d electrons is important because there are very little covalent interaction and tendency toward extension between Zn2+ and its legends. Thus, introduction of zinc into nickel obstructs the growth of particle, and hence the particle size is reduced. Further, the smaller particle sizes of the samples doped with zinc ions are due to lower bond energy of Zn2+ O2 as compared to that of N2+ O2 [37]. The reduction in particle size of a ferrite material produces interesting changes in ionic distributions of the spinel structure, which in turn can give rise to enhanced magnetic properties. 3.3 Cation Distribution Studies Distribution of cations among the available tetrahedral (A) and octahedral [B] sites of the synthesized NiZn ferrite spinels is estimated by using X-ray diffraction intensity calculations as suggested by Buerger [38], and the required formulae reproduced here are taken from [39]: Ihkl = |Fhkl |2 P Lp , (6)

planes, namely (220), (400), (422), and (440), are considered, as these planes are known to be cation-sensitive planes [40]. Instead of arbitrary intensities of individual planes (hkl), the ratios (I440 /I422 ), (I220 /I440 ), (I422 /I400 ), and (I220 /I400 ) are considered to minimize the errors in the results. The intensity ratios (I220 /I400 and I422 /I400 ) are found to be more sensitive to the cation distribution [41]. The absorption and temperature factors are not taken into account in our calculations because these do not affect the relative intensity calculations for spinels at room temperature [42]. The formulae for the structure factors for the plane (hkl) are taken as reported by Furuhashi et al. [14]. The multiplicity factors are taken from the literature [24]. The results of X-ray intensity calculations for various possible models giving rise to normal to inverse spinels have been tried and were compared with those observed intensity ratios calculated separately for entire range of zinc substitution from the X-ray diffraction data. The cation distribution for which the observed ratio agrees well with that of the proposed intensity ratios is taken as a correct one. The cation distribution proposed for the present nickelzinc nanoferrites synthesized by solgel auto combustion method is presented in Table 3. Using the accepted cation distribution, the tetrahedral site radius (rA(cal) ) and octahedral site radius (rB(cal) ) are calculated by using the relations rA(cal) = |CNi rNi2+ + CZn rZn2+ + CFe rFe3+ |, rB(cal) = 1/2|CNi rNi2+ + CZn rZn2+ + CFe rFe3+ |. (8) (9)

Theoretical lattice constant (ath ) is calculated by using the formula given by [43]: 8 ath = (rA(cal) + r0 ) + 3(rB(cal) + r0 ). 3 3 (10)

where Ihkl is the relative integrated intensity, Fhkl is the structure factor, P is the multiplicity factor for the plane (hkl), and LP is the Lorentz polarization factor, Lp = (1 + cos2 ) (sin2 cos ) . (7)

To calculate the X-ray diffraction intensities (Ihkl ) instead of considering all the planes of the spinel, typical

The obtained values of site radii and theoretical lattice constant ath are presented in Table 2. From Tables 1 and 2 it is seen that the values of observed and theoretical lattice constants are in agreement with each other favoring the estimated cation distribution [44]. Thus, it conrms that the

J Supercond Nov Magn Table 3 Estimated cation distribution analysis for Ni1x Znx F2 O4 spinel ferrite system X A-Site B-Site (I422 /I400 ) Obs. 0 0.2 0.4 0.6 0.8 1 (1Fe) (0.0Ni0.1Zn0.9Fe) (0.26Ni0.0Zn0.74Fe) (0.2Ni0.0Zn0.8Fe) (0.08Ni0.0Zn0.92Fe) (0.0Ni0.02Zn0.98Fe) [1Ni1Fe] [0.8Ni0.1Zn1.16Fe] [0.34Ni0.4Zn1.26Fe] [0.2Ni0.6Zn1.2Fe] [0.12Ni0.8Zn1.08Fe] [0.0Ni0.98Zn1.02Fe] 0.42 0.40 0.51 0.59 0.63 0.64 Cal. 0.42 0.45 0.45 0.44 0.41 0.40 (I220 /I400 ) Obs. 0.97 1.34 1.85 2.11 2.28 2.31 Cal. 1.20 1.26 1.26 1.21 1.14 1.11
Fe[B] Fe(A)

1.00 1.22 1.70 1.50 1.17 1.04

estimated cation distribution is correct. Figure 2 shows the variation of theoretical lattice constant ath with the zinc substitution x . As seen from Fig 2, the experimental lattice constant aobs is slightly smaller than that of theoretical lattice constant ath . This is in agreement with the reported results [45]. The small deviation between ath and aobs may be due to the presence of some ferrous ions Fe2+ (rFe 2+ = 0.078 nm) on octahedral sites with larger radii than Fe3+ (rFe 3+ = 0.0645 nm) [46]. Also, from Table 3 it is seen that the pure nickel ferrite shows the inverse spinel structure as expected with all the Ni2+ ions situated at octahedral [B] sites along with half of the Fe3+ ions and remaining half of Fe3+ ions at tetrahedral (A) sites. Strikingly, it is also seen from Table 3 that although Zn2+ ions have strong preference to tetrahedral (A) sites in bulk zinc ferrite, the Zn2+ ions have been found to be situated at octahedral [B] sites with large degree of inversion. This may be attributed to the method of synthesis and nanosize of the prepared ferrite samples. Hamdeh et al. [47] have reported similar results, where ne powders of zinc ferrite were produced by the supercritical aerogel method. This process has been shown to produce powders having large structural and chemical disorder. Secondly, portions of the powders were ball milled, causing much greater atomic disorder due to mechanical displacement of ions under high stress and shear. Battle et al. [48] have studied zinc ferrites prepared by variety of methods including conventional ceramic method and suggested the occupation of zinc ions at octahedral [B] sites. Further, low-temperature Mossbauer studies indicates a signicant amount of deviation of cation distribution from their bulk Preferences. Also, Fe3+ ions have strong preference for the tetrahedral (A) site as compared to octahedral [B] site [49]. So formation of pure nickel ferrite is most favorable as both Ni2+ and Fe3+ ions occupy their preferred sites with ease. As zinc is introduced in spite of its preference for (A) site, the Fe3+ ions may be forcing zinc ions to occupy octahedral [B] site as is evident from the present cation distribution study [28]. It is also reported that large cation redistributions/inversion parameters can be obtained only by placing ZnFe2 O4 into a nonequilibrium state [50]. It is also reported for specic composition of Ni1x Znx Fe2 O4 ferrite material

Fig. 4 Variation of zinc occupation level at octahedral [B] site with zinc content x

with x = 0.5 and normal site preferences; its conguration for bulk form can be written as (Zn0.5 Fe0.5 ) [Ni0.5 Fe1.5 ]O4 , which leads to the saturation magnetization value of nearly 70 emu/g at room temperature. However, in case of nanoparticles of these systems, the cation preferences do not hold good any more. This means that some of the Zn2+ will now occupy the octahedral sites and push back Ni2+ and Fe3+ to the tetrahedral sites and that the new conguration can be written as (Zn0.5x Fe0.5y Niz ) [Ni0.5z Znx Fe1.5y ]O4 such that x = y + z. Such a substitution definitely leads to the lower value of the saturation magnetization as 56.01 emu/g. A relatively low value of saturation magnetization for the sample annealed at 550 C indicates that cations are surely deviated from their normal preferences [51]. Figure 4 shows the zinc concentration at octahedral [B] site with its substitution level in nickel ferrite.

4 Conclusion Zinc-substituted nickel ferrite nanoparticlesNi1x Znx Fe2 O4 (with x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0) were synthesized successfully by using solgel auto combustion technique

J Supercond Nov Magn

with average particle size of 30 nm. The substitution of Zn2+ for Ni2+ ions results in the increasing the lattice constant. The particle size is found to decrease Zn2+ ions. The cation distribution data suggests the effect of particle size. Normally, Zn2+ ions occupy tetrahedral (A) site, but here it is observed that Zn2+ ions occupy octahedral [B] site, and it may be due to the nanosize nature of samples.

References
1. Nikumbh, A.K., Nagawade, A.V., Gugale, G.S., Chaskar, M.G., Bakare, P.P.: J. Mater. Sci. 37, 3 (2002) 2. Sickafus, K.E., Wills, J.M., Grimes, N.W.: J. Am. Ceram. Soc. 82, 12 (1999) 3. Wei, Q.M., Li, J.B., Chen, Y.J.: J. Mater. Sci. 36, 21 (2001) 4. Whinfery, C.G., Eckort, D.W., Tauber, A.: J. Am. Chem. Soc. 82, 11 (1960) 5. Raghavender, A.T., Kulkarni, R.G., Jadhav, K.M.: Chin. J. Phys. 48, 4 (2010) 6. Sawant, S.R., Patil, R.N.: Ind. J. Pure Appl. Phys. 21 (1983) 7. Cheng, S.L., Lin, J.G., Kuo, K.M., Chern, G.: J. Appl. Phys. 111, 7 07A321 (2012) 8. Singh, V.K., Khatri, N.K., Lokanathan, S.: Pramana 16, 4 (1981) 9. Wu, C.C., Mason, T.O.: J. Am. Ceram. Soc. 64, 9 (1981) 10. Tang, G.D., Ji, D.H., Yao, Y.X., Liu, S.P., Li, Z.Z., Qi, W.H., Han, Q.J., Hou, X., Hou, D.L.: Appl. Phys. Lett. 98, 072511 (2011) 11. Rietveld, H.: J. Appl. Crystallogr. 2 (1969) 12. Young, R.A.: The Rietveld Method. Oxford University Press, Oxford (1993) 13. Gonzalez, J.M.R., Arean, C.O.: J. Chem. Soc. 2155 (1985) 14. Furuhashi, H., Inagaki, M., Naka, S.: J. Inorg. Chem. 35 (1973) 15. Jadhav, K.M., Kawade, V.B., Modi, R.B., Bichile, G.K., Kulkarni, R.G.: Physica B, Condens. Matter 291(34) (2000) 16. Modi, K.B., Rangolia, M.K., Chhantbar, M.C., Joshi, H.H.: J. Mater. Sci. 41(22) (2006) 17. Ying, Y., Kim, J.M., Lee, Y.P., Kang, J.H.: J. Kor. Phys. Soc. 58(4) (2011) 18. Kadam, G.B., Shelke, S.B., Jadhav, K.M.: J. Elect. Electron. Eng. 1(1) (2010) 19. Verma, A., Dube, D.C.: J. Am. Ceram. Soc. 88(3) (2005) 20. Deka, S., Joy, P.A.: Mater. Chem. Phys. 100(1) (2006) 21. Naidu, V., Ahmed, S.K., Suganthi, M.: Int. J. Comput. Appl. 24(2), 09758887 (2011)

22. Shinde, A.B., Dhage, V.N., Jadhav, K.M.: Int. J. Eng. Adv. Technol. 2(3) (2013) 23. Zhang, R., Huang, J., Zhao, J., Sun, Zh., Wang, Y.: Energy Fuels 21 (2007) 24. Cullity, B.D.: Elements of X-Ray Diffraction 3rd edn, pp. 301 304. Chapman & Hall, London (2000) 25. Hankare, P.P., Patil, R.P., Sankpal, U.B., Jadhav, S.D., Mulla, I.S., Jadhav, K.M., Chougule, B.K.: J. Magn. Magn. Mater. 321(19) (2009) 26. Smit, J., Wijn, H.P.J.: Ferrites, p. 233. Wiley, New York (1959) 27. Kumar, P., Mishra, P., Sahu, S.K.: Int. J. Sci. Eng. Res. 2(8) (2011) 28. Upadhyay, Ch., Verma, H.C.: J. Appl. Phys. 95(10) (2004) 29. Leung, L.K., Evans, B.J., Morrish, A.H.: Phys. Rev. B 8(29) (1973) 30. Dung, N.K., Tuan, N.H.: VNU J. Sci. Math.-Phys. 25 (2009) 31. Ghazanfar, U., Siddiqi, S.A., Abbas, G.: Mater. Sci. Eng., B 118(13) (2005) 32. Haque, M.M., Huq, M., Hakim, M.A.: Ind. J. Phys. 78A(3) (2004) 33. Sathishkumar, G., Venkataraju, Ch., Sivakumar, K.: Mater. Sci. Appl. 1 (2010) 34. Attia, S.M.: Egyp. J. Solids 29(2) (2006) 35. Pandit, A.A., Shitre, A.R., Shengule, D.R., Jadhav, K.M.: J. Mater. Sci. 40(2) (2005) 36. Shannon, R.D.: Acta Crystallogr. A, Found Crystallogr. 32 (1976) 37. Shari, I., Shokrollahi, H.: J. Magn. Magn. Mater. 324 (2012) 38. Buerger, M.G.: Crystal Structure Analysis. Wiley, New York (1960) 39. Birajdar, D.S., Devatwal, U.N., Jadhav, K.M.: J. Mater. Sci. 37 (2002) 40. Ohinishi, H., Teranishi, T.: J. Phys. Soc. Jpn. 16 (1961) 41. Pandit, A.A., More, S.S., Dorik, R.G., Jadhav, K.M.: Bull. Mater. Sci. 26(5) (2003) 42. Porta, P., Stone, F.S., Turner, R.G.: J. Solid State Chem. 11 (1974) 43. Kadam, G.B., Shelke, S.B., Jadhav, K.M.: J. Electr. Electron. Eng. 1(1) (2010) 44. Sattar, A.A.: J. Mater. Sci. 39 (2004) 45. Ateia, E.: Egypt J. Solids 29(2) (2006) 46. Sebastian, M.T.: Dielectric Materials for Wireless Communication, 1st edn. Elsevier, Amsterdam (2008) 47. Hamdeh, H.H., Ho, J.C., Oliver, S.A., Willey, R.J., Kramer, J., Chen, Y.Y., Lin, S.H., Yao, Y.D., Daturi, M., Busca, G.: IEEE Trans. Magn. 31(6) (1995) 48. Battle, J., Clark, T., Evans, B.J.: J. Phys., IV 7 (1997) 49. Navrotsky, A., Kleppa, O.J.: J. Inorg. Nucl. Chem. 30, 479 (1968) 50. Deraz, N.M., Alari, A.: Int. J. Electrochem. Sci. 7 (2012) 51. Singh, R.K., Upadhyay, Ch., Layek, S., Yadav, A.: Int. J. Eng. Sci. Technol. 2(8) (2010)

Potrebbero piacerti anche