10.8
CiteScore
 
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
10.8
CiteScore
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original article
12 (
4
); 489-502
doi:
10.1016/j.arabjc.2017.02.006

Influence of the RE2O3 (RE = Y, Gd) and CaO nanoadditives on the electromagnetic properties of nanocrystalline Co0.2Ni0.3Zn0.5Fe2O4

Université de Carthage, Faculté des Sciences de Bizerte, UR11ES30, Synthèse et Structures de Nanomatériaux, 7021 Jarzouna, Tunisia
Université de Carthage, Faculté des Sciences de Bizerte, Laboratoire de Physique des Matériaux, 7021 Jarzouna, Tunisia
Université de Toulouse, LPCNO, INSA CNRS UMR 5215, 135 av. de Rangueil, 31077 Toulouse Cedex 4, France
Université Tunis El Manar, Faculté des Sciences de Tunis, Campus Universitaire, 2092, Tunisia
Northern Border University, Faculty of Science, Arar, Saudi Arabia

⁎Corresponding author at: Northern Border University, Chemistry Department, Faculty of Science, Arar, Saudi Arabia. bentaharlotfi@gmail.com (Lotfi Ben Tahar)

Disclaimer:
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.

Peer review under responsibility of King Saud University.

Abstract

The present work reports the influence of the nanoadditives Y2O3, Gd2O3, and CaO on the magnetic, electrical and dielectric properties of sintered nanoferrites Co0.2Ni0.3Zn0.5Fe2O4. All powders were synthesized via the polyol method. XRD analysis showed that except the nanoferrite which was obtained by in an one-post procedure, subsequent calcinations of the as-produced additives were necessary to obtain nanocrystals of the desired phases. The mean particle size inferred from TEM images of the nanoadditives sintered at 1000 °C ranges from 87 nm for Y2O3 to 126 nm for CaO. IR spectroscopy provided useful information on the nature of the core and the surface chemistry of the as-produced additives and their associated annealed powders. Upon sintering, it was found that the incorporation of 5 wt.% additives remarkably increased the densification of the doped materials. The most important increase in densification was observed with CaO due to its larger particles. dc M-H hysteresis loops taken at 300 K revealed a superparamagnetic behavior of the produced ferrite/nanoadditives. Additionally, as expected, the ferrite/nanoadditives showed reasonable saturation magnetization and high Curie temperature. The electrical and dielectric properties, namely the resistivity, the loss factor, and the relation frequency were found to be clearly affected by doping. The resistivity decreased with increasing temperature indicating a semiconducting behavior. Further, at room temperature, the highest resistivity was observed with Y2O3. The major role was attributed to the high fraction of insulating Y2O3 owing to its smallest particles combined with the low Fe2+ concentration in the ferrite nanoparticles taking advantages of the moderate sintering temperature. In addition, the dc conductivity was found to follow the Arrhenius law with a slope change observed at the Curie temperature. Further, all the additives clearly affected the ac conductivities of the pure ferrite. The variation of the dielectric permittivity with frequency and temperature was explained on the basis of M-W type of interfacial polarization. Additionally, at high frequencies, the lower dielectric loss was found with Y2O3 doping. It was found to be of about 10 times lower than the undoped material and much larger than reported for similar undoped bulk ferrites.

Keywords

Ferrite
Additives
Nanoparticles
Characterization
Electrical
Magnetic
1

1 Introduction

Ni–Zn spinel-type ferrites continue to be the magnetic materials of choice for high frequency technological applications over a wide frequency range (Goldman, 2006). This is due to their reasonable processing cost, excellent chemical and thermal stabilities, mechanical hardness, inherent high resistivity and high magnetic permeability, low eddy current, and low electromagnetic losses. The electromagnetic characteristics of the Ni–Zn ferrites can be tailored by judicious controlling several parameters including the nature of substituting cations, their fraction, their oxidation state and their occupancy among the spinel-type sites, the nature of the additives and their fraction, the microstructure (porosity, grain boundaries, grain size) of the final material, etc. (Goldman, 2006; Qinghui et al., 2012; Abraham, 1994; Beauclair et al., 1998; Huili et al., 2014, 2015b; Shokrollahi and Janghorban, 2007). A number of the above cited parameters are in turn highly sensitive to the microstructure of the starting materials, the type of processing technique sintering, and the associated sintering parameters (Goldman, 2006; Huili et al., 2015b). Recently, we have reported details on the electromagnetic properties of the stoichiometric nanoferrites CoxNi0.5−xZn0.5Fe2O4 (x = 0.0; 0.2; 0.4) sintered at 700 °C under 5 tons cm−2. We have demonstrated that the composition x = 0.2 should be the best promising candidate for high frequency technologies (Huili et al., 2014). Indeed, the room resistivity and the dielectric loss of the selected Co composition motioned above were respectively found to be of about two orders of magnitude larger and about an order of magnitude smaller than those measured for the unsubstituted Ni-Zn ferrite counterpart (x = 0.0). The electromagnetic properties of the ferrite Co0.2Ni0.3Zn0.5Fe2O4 were then investigated by studying the influence of the sintering conditions (the pressure and the temperature) (Huili et al., 2015b). The best electromagnetic features (low loss factor (tan δ), high resistivity (ρ), and high saturation magnetization (Msat)) were observed for the ferrite sintered in air at 1000 °C under a pressure of 7 tons cm−2. Moreover, the selected Co substituted Ni-Zn nanoferrite exhibited superior electromagnetic properties compared to similar ferrites produced by other processing, in particular those produced by the classical ceramic process. The improvement was attributed to the high quality powders in terms of microstructure and chemical homogeneity of our starting polyol-made nanoferrite. It is well established that the last processing route suffers from several drawbacks such as high energy consumption, long period of production, larger particle size, microstructural defects, the presence of impurities, the formation of non-stoichiometric ferrites, Zn and Li volatilization (Goldman, 2006; Rahaman, 2003). Electromagnetic properties of ferrites can be remarkably improved by adding an optimized small amount of oxide (including their combination) such as CaO, Bi2O3, Na2O, SiO2 and MnO2. Further, in combination with adjustable sintering conditions, doping enables obtaining high-performance ferrites (Goldman, 2006; Tasaki and Izushi, 1977; Rezlescu et al., 2000; Mirzaee, 2014; Rao et al., 2004). The advantages reside in the main fact that densification can be largely increased by adjusting the above cited parameters, namely the doping content and the sintering conditions. Further, densification improvement coupled with high electrical resistivity could offer better miniaturization of the electronic components (Lebourgeois et al., 1992). Additionally, energy dissipation at high frequencies is substantially reduced in fine-grained monodomain Ni-Zn ferrites, which is attributed to the absence of intragranular domain walls (Van der Zaag et al., 1996).

The present work focuses on studying the influence of 5 wt.% of three nanoadditives (Y2O3, Gd2O3, and CaO) on the microstructure and the magnetic and the electrical/dielectric properties of sintered nanopowders of Co0.2Ni0.3Zn0.5Fe2O4. It aims at further improving the electromagnetic properties of the sintered nanoferrite recently reported by us (Huili et al., 2015b). Nanocrystalline powders of the oxides were successfully produced by the polyol method and subsequent sintering at moderate temperatures. The structure and the microstructure characterization of both as-produced powders and their associated annealed ones at 1000 °C were achieved by XRD, IR spectroscopy, and TEM.

2

2 Experimental procedure

2.1

2.1 Synthesis

Co0.2Ni0.3Zn0.5Fe2O4 nanoparticles were produced by the polyol method starting from Fe (III) chloride and the acetates of M(II) (M = Co, Ni, and Zn) as precursors and diethyleneglycol as a solvent (Huili et al., 2015a). Stoichiometric nanoparticles were obtained by optimizing four synthetic parameters including the hydrolysis ratio (h), the acetate ratio (τAc), and the Zn ratio (τZn), defined by the molar ratios of water/metal, acetate/M2+, and Zn2+/Fe3+, respectively, and the refluxing duration (t). The values of the four parameters h, τAc, τZn, and t, were found to be ≈1, 3.66, 0.5, and 3 h, respectively. The above cited optimized synthetic parameter values found for Co0.2Ni0.3Zn0.5Fe2O4 were adopted for the preparation essays of Y2O3, Gd2O3, and CaO powders using the polyol method. The respective precursors of the oxides are Y(CH3CO2)3 (Sigma-Aldrich; 99.9%), Gd(CH3CO2)3 (Sigma-Aldrich; 99.9%), and CaCO3 (Sigma-Aldrich; 98.5%). All chemicals were used as received. In a typical experimental procedure, an appropriate amount of each precursor and 62.5 mL of diethyleneglycol were added to a round bottom three-necked flask 250 mL capacity to reach a nominal concentration of the associated cation of 0.2 M. Then, appropriate amounts of distilled water and anhydrous sodium acetate (Sigma-Aldrich; 99%) were added to adjust the hydrolysis and the acetate molar ratios to the values ≈1 and 3.66, respectively. The resulting mixture was then refluxed at boiling temperature under mechanical stirring for 3 h. After slow cooling, the obtained precipitate was washed with ethanol under vigorous sonication with intermittent centrifugation. Finally the precipitate was dried overnight in air at 60 °C. Importantly, we also note that unlike the ferrite phase which was obtained without any further heat treatment, formation of crystalline Y2O3, Gd2O3 and CaO required subsequent calcinations of their associated as-produced powders (the dried precipitates). Further details on the calcination conditions will be given in the results and discussion section.

2.2

2.2 Characterization techniques

The structure and phase purity of the produced powders were examined by X-ray diffraction (XRD) using a BRUKER D8 ADVANCE diffractometer equipped with a copper anode (λCuKα = 1.540/1.544 Å). XRD powder pattern indexing was carried out with McMaille program (Le Bail, 2004) and the unit cell parameters were refined by the Rietveld method (Rietveld, 1969) using the FULLPROF program (Carvajal, 1998). The average crystallite size was inferred from the broadening of the XRD diffraction peaks applying the Scherrer formula (Scherrer, 1918). Instrumental broadening was estimated from Cagliotti plots (Cagliotti et al., 1958) using a polycrystalline silicon as standard. IR characterization was carried out on a Nicolet UR 200 FTIR spectrometer using the ATR technique over the range 400–4000 cm−1 with a resolution of 4 cm−1. The particle morphology was observed by a FEI QUANTA 200 transmission electron microscope (TEM) operating at 200 kV with a probe diameter of 200 nm. A small quantity of the powder was ultrasonically dispersed in ethanol, then a drop of the suspension was deposited on a carbon coated copper grid and the solvent was evaporated at room temperature. TEM images were recorded at different magnifications at randomly selected areas of sample and the average particle size was obtained by measurements of at least 100 particles. dc magnetic measurements were conducted on a Quantum Design PPMS VSM magnetometer. The sintered powders were compacted inside capsules to prevent physical movement during the experiments. The hysteresis loops were measured with magnetic field cycling in the ±30 kOe range at 300 K. The temperature dependence of the spontaneous magnetization was recorded between 50 and 330 K under a constant magnetic field of 20 kOe. The collected data were corrected from the diamagnetic contribution and were presented in the cgs–emu g−1 units. The collected data were corrected from the diamagnetic contribution of the sample holder and were presented in cgs–emu g−1 units. The dielectric response of the sintered pellets was measured in the heating mode from 105 °C to 400 °C using TEGAM 6192 ALF impedance analyzer in the frequency range 10 Hz–13 MHz. Electrical equivalent circuits of the sintered pellet were simulated using Zview program (Johnson, 2005). The as-produced ferrite and the as-produced additives (Y2O3, Gd2O3 or CaO) were individually pre-annealed in air at 500 °C for 4 h in order to remove the organic species grafted onto the nanoparticles surface (Huili et al., 2015a). Then, an appropriate mass of the resulting ferrite powder was then mixed with a small quantity (5 wt.%) of each additive. Each mixture was then well ground using agate mortar and then uniaxially compacted under 7 tons cm−2 pressure resulting in disk–shaped pellet of about 13.0 mm diameter and 1.5 mm thickness. The pelletized powders were then heated in air for two hrs at 1000 °C in a programmable muffle furnace followed by a slow cooling to room temperature. Hereafter, for the sake of simplicity, the sintered pellets will be sometimes denoted as undoped ferrite, ferrite/Y2O3, etc., for the sintered pure ferrite Co0.2Ni0.3Zn0.5Fe2O4 and the sintered 5 wt.% Y2O3 dopped ferrite, etc., respectively. Finally, the surfaces of the pellets were coated by silver paste in order to ensure a good and uniform contact with the incoming current wires.

3

3 Results and discussion

3.1

3.1 Phase and microstructure analyses

The phase and microstructure analyses of the as-prepared ferrite and its corresponding annealed one at 1000 °C were reported elsewhere (Huili et al., 2015b). The XRD patterns are indexed in a spinel-type structure with no evidence of any foreign phase, indicating, in particular, a high thermal stability of the sintered ferrite. Crystallites of both powders were found to be in the nanometer scale. The mean crystallite size in the as-produced ferrite and its associated sintered powder were found be at around 2 nm and 86 nm, respectively. Figs. 1–3 show the XRD patterns of the as-synthesized powders of the pure additives and their associated ones calcinated at different conditions as well as those of ferrite/additives powders sintered at 1000 °C. The JCPDS Bragg positions of Y2O3, Gd2O3, and CaO standards are also plotted for comparison. One can observe that for the as-prepared powders of Y2O3 and Gd2O3, the diffraction patterns are poorly resolved, indicating the possibility the formation of amorphous intermediate phase and/or ultrasmall nanoparticles. At the annealing temperature 500 °C, appears the signature of both phases Y2O3 (JCPDS card No. 088-1040) and Gd2O3 (JCPDS card No. 043-1014) with broad XRD peaks indicating the formation of ultrafine crystallites. With the increase in annealing temperature, the diffraction peaks are expected to become more intense and sharper in accordance with an increase in crystallinity and crystallite size. The mean crystallite size (〈LXRD〉) of Y2O3 and Gd2O3 annealed at 1000 °C was found to be at around 81 nm and 108 nm, respectively. Table 1 gathers selected structural and structural parameters of the produced powder oxides annealed at 1000 °C. Additionally, at 1000 °C all the observed peaks of the produced RE2O3 (RE = Y, Gd) were assigned to the cubic structure of the reference oxides without any detectable extra phase indicating in particular highly thermally stable oxides. The refined lattice parameters (Table 1) of RE2O3 (RE = Y, Gd) annealed at 1000 °C were found to be in excellent agreement with those reported elsewhere (Baldinozzi et al., 1998; Zachariasen, 1927). Fig. 3 shows the XRD patterns of the as-prepared CaO and the associated annealed powder at 1000 °C along with standard JCPDS data of calcium oxide and calcium carbonate (calcite allotrope). It can be observed that the as-prepared powder consists of pure calcite (JCPDS card No. 5-586) indicating the chemical inertness of the major phase precursor (calcite) toward the polyol solvent. Additionally, a close look of the XRD pattern of the as-produced revealed the disappearance of an impurity initially found with the precursor. We also note that a number of attempts varying the hydrolysis ratio, the acetate ratio, and the refluxing time remained unsuccessful for the preparation of the expected oxide. On the other hand, the annealed powder at 1000 °C as expected reveals the formation of pure calcium oxide (JCPDS card No. 78-0649) through a well-known thermal decomposition of calcium carbonate. In CaO powder the mean crystallite size is notably larger than in the other additives, and it is of about 270 nm. The XRD patterns of the sintered ferrite/additives show the signature of the two components of each powder indicating in particular the occurrence of neither solid solution (there’s no shift of diffraction peaks) nor detectable other new phases i.e. each component of the sintered materials retains its chemical and structural identity. The relatively moderate sintering conditions (1000 °C, 2 h) compared to the conventionally adopted for the sintering of bulk ferrites (>1200 °C, >12 h) could be the main reason for the absence of any detectable solid-state reaction between the additives and the ferrite phase. However, the co-sintering of the ferrite and the additives phases may affect their microstructure in comparison with their own microstructure. The microstructural features of each component of the ferrite/additives powders were estimated (Table 1). Except for the CaO phase whose size was surprisingly found to be reduced to about the half, the mean crystallite size and the strain of the component phases of each sintered powder are found to be similar to those calculated for the isolated phases. In order to get more information about the powder morphology of the additives Y2O3, Gd2O3, and CaO, TEM analysis was performed (Fig. 4(a–c)). As can be clearly seen from TEM images and the associated histograms, the grains of RE2O3 (RE = Y, Gd) show important irregularity in the size and the shape with a high degree of agglomeration. Those of Y2O3 are relatively more regular with polygonal plate-shaped nanoparticles of six or higher sides and a mean particle size (the average size for each particle was calculated by measuring the largest and the smallest length) of about 87 nm. On the other hand CaO powder consists of roughly spherical, slightly agglomerated nanoparticles with a particle size ranging from about 90 to 180 nm and an average diameter of 126 nm. The FT-IR spectroscopy was used to elucidate the surface functionality and the vibrational modes of the inorganic core of the as-prepared additives (Fig. 5). The IR spectra of the solvent diethyleneglycol and as-prepared Y2O3 annealed at 1000 °C are also recorded for comparison. As expected the IR spectrum of the as-prepared CaO powder shows a very strong absorption band at around 1430 cm−1 and two sharp ones centered at 1430, 880, and 713 cm−1 arising from the stretching modes of carbonate ion (Miller and Wilkins, 1952). For the as-prepared powders of RE2O3 (RE = Y, Gd), the IR spectra in the region above 750 cm−1 show characteristics of the main chemisorbed species including hydroxyl groups, polyol, and acetate ions (Huili et al., 2015a). For instance, we noticed the presence of the alkyl-substituted ether, C–O–C, bridging and secondary alcohol, C–O, stretching bands (1150–1050 cm−1) of the polyol as well as the characteristic acetate, COO, symmetric νs (weak and broad at 1407–1440 cm−1) and asymmetric νas (strong and broad at 1622–1650 cm−1) stretching bands. These assignments are supported by the presence of a couple of sharp bands located at 2920/2880 cm−1 associated with C-H vibrations (νass CH2). It is worth noting that the organic species detected on RE2O3 particles are completely absent in the case of the as-prepared CaO sample thus supporting the XRD results. Another important point is to be noted that the IR bands associated with the capping agents are much more pronounced with the as-prepared Y2O3 than the observed for as-prepared Gd2O3. The result can be explained on the basis of the difference in the specific surface of the particles of the two samples; the smaller the particle size, the larger the surface area is, and therefore the larger the amount of grafted organic species on its surface is. The last result supports the XRD microstructure analyses results found for the two sintered samples. For the annealed two oxides at elevated temperatures the organic species are expected todepart from the particles surface through a combustion reaction. The result can be clearly seen for the typical IR spectrum of the as-prepared Y2O3 powder annealed at 1000 °C which revealed the complete disappearance of the IR features of the organic. Additionally, for both the as-prepared Gd2O3 and the as-prepared Y2O3 powder annealed at 1000 °C appeared at low frequencies two broad intense bands in the range 600–400 cm−1. The absorption bands are characteristic of the rare earth oxides RE2O3 (including Y2O3) cubic oxides (Baun and Mc Devitt, 1963). Further, as can be deduced from IR study, unlike the as-prepared Y2O3, appearance of characteristic bands for the as-prepared Gd2O3 in the range 600–400 cm−1 confirms the possibility of one-pot synthesis of the last oxide without any further heat treatment. However, as pointed out above, XRD demonstrated that the Gd2O3 core is of amorphous nature since no diffraction peaks were detected and subsequent annealing is just required for promoting its crystallization. Furthermore, it is interesting to note that the vibrational frequency (≈533 cm−1) observed for sintered Y2O3 is notably higher than that observed for Gd2O3 (≈506 cm−1). The result can be correlated with RE-O bonding length; the oxide with the larger RE-O bond length, i.e. with the larger cations radius (Gd3+) is expected to exhibit the lower stretching frequency IR band.

XRD patterns of the as-prepared sample obtained from the preparation of Y2O3 and associated powders annealed at different conditions in comparison with the reference Y2O3 (JCPDS card No. 88-1040). The XRD peaks of the spinel-type ferrite phase (JCPDS card No. 52-0278) are indicated by the symbol (+). For the as-prepared powder, the only peak observed at around 27° indicated by (*) is due to the sample holder.
Figure 1
XRD patterns of the as-prepared sample obtained from the preparation of Y2O3 and associated powders annealed at different conditions in comparison with the reference Y2O3 (JCPDS card No. 88-1040). The XRD peaks of the spinel-type ferrite phase (JCPDS card No. 52-0278) are indicated by the symbol (+). For the as-prepared powder, the only peak observed at around 27° indicated by (*) is due to the sample holder.
XRD patterns of the as-prepared sample obtained from the preparation of Gd2O3 and the associated powders annealed at different conditions in comparison with the reference Gd2O3 (JCPDS card No. 43-1014). The XRD peaks of the spinel-type ferrite phase (JCPDS card No. 52-0278) are indicated by the symbol (+).
Figure 2
XRD patterns of the as-prepared sample obtained from the preparation of Gd2O3 and the associated powders annealed at different conditions in comparison with the reference Gd2O3 (JCPDS card No. 43-1014). The XRD peaks of the spinel-type ferrite phase (JCPDS card No. 52-0278) are indicated by the symbol (+).
XRD patterns of the as-prepared sample obtained from the preparation of CaO and the associated powder annealed at 1000 °C in comparison with the references calcite (JCPDS card No. 05-0586) and calcium oxide (JCPDS card No. 017-0912). (*) unassigned peak due to a minor unknown impurity in CaCO3 precursor (Sigma-Aldrich; 98.5%). The XRD peaks of the spinel-type ferrite phase (JCPDS card No. 52-0278) are indicated by the symbol (+).
Figure 3
XRD patterns of the as-prepared sample obtained from the preparation of CaO and the associated powder annealed at 1000 °C in comparison with the references calcite (JCPDS card No. 05-0586) and calcium oxide (JCPDS card No. 017-0912). (*) unassigned peak due to a minor unknown impurity in CaCO3 precursor (Sigma-Aldrich; 98.5%). The XRD peaks of the spinel-type ferrite phase (JCPDS card No. 52-0278) are indicated by the symbol (+).
Table 1 Selected structural and microstructural characteristics inferred from XRD and TEM for Y2O3, Gd2O3, and CaO powders obtained by annealing the associated as-produced powders at 1000 °C. Data of Co0.2Ni0.3Zn0.5Fe2O4 were reported by us in a previous work (Huili et al., 2015b).
Lattice parameter/Å 〈LXRD〉/nm Strain/% 〈DTEM〉/nm
Co0.2Ni0.3Zn0.5Fe2O4 8.403(4) 145.0 0.14 86 ± 21
Y2O3 10.574(3) 81.2 0.22 87 ± 19
Gd2O3 10.803(3) 108.1 0.26 112 ± 23
CaO 4.751(3) 268.7 0.09 126 ± 14
Ferrite/Y2O3 154.1/95.5 0.13/0.22
Ferrite/Gd2O3 156.5/102.5 0.13/0.22
Ferrite/CaO 149.4/124.1 0.16/0.15
TEM images for Y2O3 (a), Gd2O3 (b), and CaO (c), sintered at 1000 °C with their corresponding particle size distribution histograms.
Figure 4
TEM images for Y2O3 (a), Gd2O3 (b), and CaO (c), sintered at 1000 °C with their corresponding particle size distribution histograms.
IR spectra of the as-prepared produced powders. The IR spectra of the solvent diethyleneglycol and as-prepared Y2O3 annealed at 1000 °C are also recorded for comparison. The grid lines are a guide for eyes.
Figure 5
IR spectra of the as-prepared produced powders. The IR spectra of the solvent diethyleneglycol and as-prepared Y2O3 annealed at 1000 °C are also recorded for comparison. The grid lines are a guide for eyes.

3.2

3.2 Magnetic properties

The isothermal hysteresis loops of the three ferrite/additives measured at 300 K are presented in Fig. 6. As can be noted, there is no hysteresis features (neither remanence nor coercivity), indicating a superparamagnetic behavior of the materials at ambient temperature. In small ferromagnetic or ferrimagnetic single domain particles, superparamagnetism occurs above the so-called blocking temperature, TB, because of weakly interacting and thermal fluctuations of the spins of the nanoparticles. The thermal effects allow flips of spins between the easy magnetization axes which lead to near zero–coercivity and increase in saturation magnetization (Pileni, 2001). Below TB the thermal fluctuation does not dominate and the superparamagnetic nanosized particles cannot rotate freely but freeze in random orientations leading to high–coercive fields. The saturation magnetization, Msat, values gathered in Table 2 are determined by extrapolating a linear fitting of M vs 1/H data to 1/H = 0. As expected the saturation magnetization of the ferrite/additives is quite similar to each other and doesn’t differ significantly from that of the pure sintered ferrite owing to the low fraction (5%) of the additives in the materials. The slight decrease of Msat observed for the ferrite/additives in comparison with the undoped ferrite (Table 2 and the zoom-view in Fig. 6), is possibly attributed to the very weak ferromagnetism and diamagnetism at room temperature of nanosized Gd2O3 and Y2O3 (Bud’ko and Canfield, 2006; Zhu et al., 2012) and CaO, respectively. It is worth noting that dissolution of any of the Gd3+, Y3+, or Ca2+ cations in the spinel-like structure, even at very low concentration should appreciably alter the magnetization of the ferrite material. For instance, it is reported that entrance of Gd3+ and Y3+ at low concentration, respectively enhanced and reduced the saturation magnetization of CoFe2O4 (Yan et al., 1998). Therefore any assumption of incorporation of the above mentioned cations in the spinel-type lattice can be ruled out. Decay of magnetization with temperature and the Curie temperature of spinel ferrites are of vital important parameters from technological applications’ point of view. To get a quantitative description we measured the temperature dependence of the spontaneous magnetization, Msp (Fig. 7). For nanostructured ferrites, below the Curie temperature Msp vs T can be well fitted with a modified Bloch’s law function (Abdallah et al., 2014).

(1)
M sp ( T ) = M sp ( 0 ) [ 1 - ( T / T C ) β ] where Msp(0) is the spontaneous magnetization at 0 K, TC is the Curie temperature, and β is the so-called Bloch’s exponent. Msp(0) and TC values correspond to the intercept of the modified Bloch’s law curves with the Msp(T) and T axis, respectively. An excellent fitting (R2 = 0.999) with a modified Bloch’s law was obtained. The modified Bloch’s law parameters are gathered in Table 2. The Bloch’s exponent, β, ranges between 1.12 and 1.32. These values are quite smaller than the constant value of 1.5 calculated for a three-dimensional system (Cullity, 1972). In general, Bloch’s exponent depends upon the magnetic structure, the microstructure, and the surface treatment of the material (Zhang et al., 1998). In our case, the variation of β of the three ferrite/additives compared to each other’s and to that of a three dimensional system is likely to be attributed to the difference in microstructure of the materials as observed by TEM. As shown in Fig. 7, the magnetization decay rate is low which allows the materials to maintain reasonable magnetization in a large temperature range. For instance a reduction in magnetization of 50% is reached at a temperature of about 340 K (97 °C). The Curie temperature (Table 2) of all the ferrite/additives materials exceeded 300 °C. The TC values are quite high which permits the materials to work in severe conditions. In particular, the ferrite/additives can be valuable for special applications including military, engine technology which require very high operating temperatures (Goldman, 2006). For bulk materials Curie temperature depends on the A-B and B-B magnetic interaction strength which in turn is sensitive to the chemical composition and cation occupancy among the A and B spinel sites. For magnetic nonmaterials additional contribution causing TC reduction due to finite size effects has been also reported (Batlle and Labarta, 2002). Thus, for our nanoferrites, the slight difference between the transition temperatures, TC, could be attributed to the finite-size effect of the constituting nanoparticles.
Dc M-H hysteresis loops measured at 300 K for ferrite/Y2O3, ferrite/Gd2O3, and ferrite/CaO, sintered at 1000 °C. The inset is a zoom-view at high fields. The data of undoped ferrite were reported by us in a previous work (Huili et al., 2015b).
Figure 6
Dc M-H hysteresis loops measured at 300 K for ferrite/Y2O3, ferrite/Gd2O3, and ferrite/CaO, sintered at 1000 °C. The inset is a zoom-view at high fields. The data of undoped ferrite were reported by us in a previous work (Huili et al., 2015b).
Table 2 Selected magnetic characteristics of samples sintered at 1000 °C under a pressure of 7 tons cm−2 of the undoped ferrite, Co0.2Ni0.3Zn0.5Fe2O4, and their associated doped ones. Data of Co0.2Ni0.3Zn0.5Fe2O4 were reported by us in a previous work (Huili et al., 2015b).
Undoped ferrite Ferrite/Y2O3 Ferrite/Gd2O3 Ferrite/CaO
Msat (300 K)/emu g−1 75 71 73 70
Msp(0)/emu g−1 138 129 129 124
TC/K 600 600 587 575
β 1.32 1.12 1.18 1.18
Spontaneous magnetization, Msp vs temperature measured at 20 kOe for ferrite/Y2O3, ferrite/Gd2O3, and ferrite/CaO, sintered at 1000 °C. The inset is a zoom-view around 220 K showing the goodness of fit to the modified Bloch’s law.
Figure 7
Spontaneous magnetization, Msp vs temperature measured at 20 kOe for ferrite/Y2O3, ferrite/Gd2O3, and ferrite/CaO, sintered at 1000 °C. The inset is a zoom-view around 220 K showing the goodness of fit to the modified Bloch’s law.

3.3

3.3 Electrical and dielectric study

The electrical and dielectric properties of the sintered Co0.2Ni0.3Zn0.5Fe2O4 doped with 5 wt.% of Y2O3, Gd2O3, and CaO were studied by complex impedance spectroscopy as a function of temperature and frequency.

The following are useful equations for the determination of a number of electrical/dielectric quantities based on the measured real (Z′) and imaginary (Z″) components of the complex electrical impedance (Z) of the electrical system (pellet + electrode + …).

The real (ε′) and the imaginary (ε″) parts of the complex dielectric constant, ε ( ε = ε + j ε ), and applying the following relations:

(2)
ε = - Z ω C 0 Z 2 + Z 2
(3)
ε = Z ω C 0 Z 2 + Z 2

The dielectric loss factor (tan δ) is calculated from the ratio of ε″ to ε′:

(4)
tan δ = ε ε where the following symbols stand for the following:

  • f, the frequency of the applied external electrical field.

  • ω= 2πf, the angular frequency.

C 0 = ε 0 S t , the free space capacitance of a parallel-plate capacitor with separation t and face area S those of the studied pellets.

The complex electrical modulus, M, is defined as the inverse of complex permittivity:

(5)
M ( f ) = 1 ε ( f ) = M + jM

Taking account of the above cited relations, the real (M′) and the imaginary (M″) parts of the complex electrical modulus can be calculated as follows:

(6)
M = - ω C 0 Z
(7)
M = ω C 0 Z
where the meaning of the parameters ω and C0 is those defined before.

The dc conductivity, σdc, can be calculated applying the following formula:

(8)
σ dc = t SR where t, S, and R are the thickness, the cross section of the disk-shaped pellet, and the total resistance inferred from −Zvs Z′ plots, respectively.

The percentage porosity was also calculated according to the relation:

(9)
Porosity ( % ) = ρ th - ρ exp ρ th × 100 where ρth and ρexp denote the theoretical inferred from XRD data, and experimental density calculated from dimensions of the sintered pellets, respectively. Porosity values of all sintered samples are presented in Table 3. As can be seen the porosity is lowered in the presence of any of the additives, the most important decrease in the porosity was obtained with CaO. A comparable result was reported earlier by Rezlescu with a Zn-rich bulk Ni-Zn ferrite doped with Sb2O3, Na2O, CaO, and ZrO2 (Rezlescu et al., 2000). One can suggest that the larger grains prefer to fill up the porosity of the host ferrite which increases the densification, whereas small grains probably prefer to spread into grain boundaries. As pointed out before, change in porosity (i.e. change in intergranular contact and space) can significantly affect the electromagnetic properties of the studied ferrites, such the resistivity, the loss factor, and the Curie temperature. The Cole–Cole diagrams (−Zvs Z′) at selected temperatures of the samples under investigation are depicted in Fig. 8(a-d). For the two samples Co0.2Ni0.3Zn0.5Fe2O4/Gd2O3 and Co0.2Ni0.3Zn0.5Fe2O4/CaO (Fig. 8(c) and (d)), it can be observed that impedance plots are located on two arcs. The arc on the low-frequency side (high Z′) is due to the grain boundary conduction and that on the high-frequency side (low Z′) is due to the grain conduction. The contribution from grain interior and grain boundary effects is manifested by the diameter of respective semicircular arcs. Analysis of the electrical properties was made by assuming the equivalent circuits shown in the inset of Fig. 8(c and d). The characteristics (Table 3) of the equivalent circuits were obtained by fitting to the experimentally observed impedance data. The adequate equivalent circuits consisted of a series combination of Rs + (R1//CPE1) + (R2//CPE2), where the symbols R and CPE represent a resistance and a constant phase element, respectively (Huili et al., 2015b). The element Rs is attributed to the contact resistance electrode-pellet response. The second part of the circuit consists of parallel combination of a resistance R1 and a CPE1 (R1//CPE1) attributed to the grains, while the third portion (R2//CPE2) is ascribed to the grains boundary. For the undoped ferrite Co0.2Ni0.3Zn0.5Fe2O4 and the doped ferrite Co0.2Ni0.3Zn0.5Fe2O4/Y2O3, however, only a single semicircular arc is present (Fig. 8(a and b)) indicating a negligible grain boundary contribution. Their associated equivalent circuits are therefore composed of only a series combination of Rs + (R1//CPE1). The total resistance, R, was determined from the intercept of the Z′–axis. From Table 3 one can see that for all the studied samples, the resistance clearly decreased with temperature increase typical of a semiconductor behavior. The temperature dependence of the total dc conductivity measured up to a temperature of about 400 °C is shown in Fig. 9. It can be visualized from the figure that the experimental points of Ln (σdc·T) vs 1000/T showed an almost linear variation with a slope change at a critical temperature. Comparable trends have been already observed with a number of mixed ferrites such as Ni-Zn and Li ferrites (Murthy and Sobhanadri, 1976; Ravinder, 1991). The critical temperature at which occurred a slope change is the Curie temperature, TC which is the temperature above which the spontaneous magnetization of a ferromagnetic or ferrimagnetic material vanishes. A change in magnetic ordering occurs from ferro- or ferrimagnetism to paramagnetism (Ravinder, 1991; Smit and Wijn, 1959). The estimated values are depicted in Table 4. In both the ferrimagnetic and the paramagnetic regions the experimental values of the conductivity can be fitted by an equation according to the Arrhenius law:
(10)
σ dc . T = A exp ( - E a / K B T )
where A is a temperature independent quantity, Ea is the activation energy needed to release an electron for hopping from an ion to the next, KB is the Boltzmann constant, and T is the temperature in Kelvin. In spinel ferrites, the conductivity mechanism is caused by electron hopping between Fe3+ and Fe2+ cations at octahedral sites (Goldman, 2006; Verway and Haayman, 1941). Conduction magnitude in ferrites is actually a result of the combined influence of several factors (more or less interdependent) such as the concentration of Fe3+/Fe2+ pairs present at octahedral sites, the density, the grain size, the type of additives, and the sintering history. Among these parameters Fe3+/Fe2+ pair concentration and the mobility of the ions are reported to play the dominant role in the process of conduction as well as in dielectric polarization (Goldman, 2006; Smit and Wijn, 1959). It is observed from Fig. 9, that for a given sample, the dc conductivity increased with temperature. This increase is mainly due to the increase in mobility of the charge carriers which is reported to be highly temperature-dependent quantity, while the carrier concentration is claimed to be almost unaffected by temperature variation (Ravinder, 1991; Smit and Wijn, 1959). Moreover, at elevated temperatures as high as ≈127 °C (400 K), addition of Y2O3, Gd2O3, and CaO led to more conductive materials. The highest conductivity obtained for Co0.2Ni0.3Zn0.5Fe2O4/CaO is likely to be linked to the significant decrease in porosity (from ∼18 for the undoped ferrite to ∼6% for the CaO doped one) (Table 3). On the other hand, regarding conductivity of Gd2O3 and Y2O3 doped ferrites at low temperatures below a threshold value (127 °C for the Gd2O3 and 102 °C for Y2O3) an inversion of conductivity can be observed; the conductivity of the materials become lower than that of the undoped ferrite. This may be caused by a complex variation of the mobility as a function of temperature. Extrapolation from the Arrhenius plots of the electric resistivity (the inverse of the conductivity) values at room temperature (278 K, 1000/T = 3.35) is listed in Table 4. These values were remarkably higher than those reported for bulk doped and undoped Ni-Zn ferrites with comparable chemical composition (Smit and Wijn, 1959; Mohan et al., 1999; Shahjahan et al., 2014). In our case, the negligible amount of Fe2+ and the installation of a finite size-dependent metastable cation’s occupancy over the spinel-type sites due to low sintering conditions should be the major factors inducing high resistivity of the nanomaterials in comparison with the bulk ferrites (Huili et al., 2014). It has been reported that formation of the ferrous ions Fe2+ is consequence of Zn volatilization during sintering (Rahaman, 2007) which obviously increases with increase in sintering temperature. Further the larger resistivity was found with Y2O3 doping. It is of about two orders of magnitude larger than that measured for the undoped ferrite, namely Co0.2Ni0.3Zn0.5Fe2O4. The highest resistivity observed with Y2O3 doping correlated its relatively smaller particles (Shokrollahi and Janghorban, 2007); Small grains imply larger number of insulating grain and hence greater energy barriers to electron conduction resulting thereby in higher resistivity. The activation energies in the ferrimagnetic (Ef) and paramagnetic (Ep) regions are calculated from the slopes of Arrhenius plots. An examination of Table 4 reveals two conclusions: (i) for each of the four samples under study, Ep value is larger than the Ef value. Similar results have been reported elsewhere (Ravinder, 1991). (ii) Despite the decrease in porosity on doping, the activation energy of the doped samples is higher than that of the undoped one. This can be explained by the fact that introduction of the doping phase in the host material induces heterogeneous areas of grain to grain contact which reduces the probably transfer of electrons between grains of the ferrite phase. Frequency dependence plots of the ac electrical conductivity, σac, of Co0.2Ni0.3Zn0.5Fe2O4/additives at selected temperatures are presented in Fig. 10. For a fixed temperature, the conductivity is almost constant over a wide range of frequency. Beyond a temperature–dependant frequency threshold, however a clear dispersion with frequency is observed. Besides, it can be seen the ac conductivity increases with temperature indicating a semiconducting behavior of the studied materials. Variation of ac conductivity with temperature and frequency can be expressed by Jonsher’s universal power law (Jonsher’s, 1974):
(11)
σ ac = σ 0 ( T ) + B ( T ) ω n ( T )
where the first term σ 0 represents the dc conductivity and it is only temperature-dependent. The second term is a frequency and temperature-dependent function in which B is the dispersion parameter and n(T) is the universal power law exponent lying in the range 0–1 for typical ac conduction through hoping mechanism. Both B and n are constants that depend upon the temperature and composition. The fractional exponent n(T) is a measure of the degree of correlation between the ac conductivity and frequency, and it should be zero for a random hopping and tends to 1 as correlation increases (Huili et al., 2015b; Singh et al., 2011). The fractional exponent, n(T), was estimated from a linear fit of Ln σac vs Lnω in the high frequency region according to Eq. (11). n(T) values were found to vary in the range 0.30–0.69, 2.42 × 10−3–6.80 × 10−3, 5.01 × 10−4–1.00 × 10−3, and 1.80 × 10−3–0.59 for the undoped ferrite, ferrite/Y2O3, ferrite/Gd2O3, and ferrite/CaO, respectively. The significant variation of the fractional exponent with temperature is a strong evidence for thermally activation of charge carriers at high frequencies. At a fixed temperature, the increase in the ac conductivity with frequency can be explained on the basis of Maxwell–Wagner (M-W) two layer model (Maxwell, 1973; Wagner, 1913; Koops, 1951); the first layer is associated with a well-conducting grains separated by layers of lower conductivity consisting of the grain boundary. The resistive grain boundaries were found to be more effective at lower frequencies while the conductive grains are more effective at higher frequencies (Patange et al., 2011). In spinel-type ferrites, the electrical conductivity is explained on the basis of electron hopping mechanism between nearest Fe2+ and Fe3+ ions located at octahedral sites (Verway and Haayman, 1941; Gul et al., 2007). On the basis of the two conducting models, it can correlate the low and high conductivity layers with a low and high hopping frequency between the two iron cations, respectively. Further, as shown in Fig. 10(a–d), at a given temperature and frequency, the additives Y2O3, Gd2O3, and CaO lead to a significant increase of σac. The higher conductivity was obtained with Co0.2Ni0.3Zn0.5Fe2O4/CaO for which the porosity is notably larger than that measured for the other materials. Further, as pointed out before the difference between conductivity observed for our nanomaterials should be correlated with the difference in both their interdependent porosity and intergrain surface contact and chemical composition at the interface. The two parameters are themselves strongly dependent on size and shape of the grains. The electrical relaxation in ionically and electronically conducting materials has been extensively studied and analyzed in terms of modulus formalism (Macedo et al., 1972). The frequency dependence of the real (M′) and the imaginary (M″) parts of the electrical modulus measured at selected temperature is shown in Figs. 11 and 12. At low frequency all curves show dielectric dispersion with an almost independent frequency and temperature variation. Beyond a temperature-dependent threshold there is a sharp increase in frequency. Comparable trends were quoted for a variety of ferrospinels (Ponpandian et al., 2002). The transition from the frequency-independent to frequency-dependent modulus components indicates the onset of a relaxation phenomenon. Moreover, M″ curves show an asymmetric peak at a frequency threshold, fr. fr corresponds to the most probable relaxation frequency for which charge carriers can move over a long distance (i.e., the charge carriers can perform successful hopping from one site to the neighboring site). This hopping frequency corresponds also to the dipolar orientational relaxation frequency of the dipoles determined by the electron exchange between Fe3+ and Fe2+ ions. It is interesting to note that the hopping frequencies for all the samples have temperature dependence similar to that of the dc conductivity. The increase in relaxation frequency with temperature is due to the thermal activation of the localized electric charge carriers forming dipoles, which are responsible for dielectric polarization (Van Uitert, 1955).
Table 3 Equivalent circuit parameters determined at selected temperatures (in °C) of samples sintered at 1000 °C under a pressure of 7 tons cm−2 of the undoped ferrite, Co0.2Ni0.3Zn0.5Fe2O4, and its associated doped ones. Data of Co0.2Ni0.3Zn0.5Fe2O4 were reported by us in a previous work (Huili et al., 2015b).
Porosity% 250 275 300 325
Undoped ferrite 18 Rs (Ω) 99 19 104 131
R1 (Ω) × 104 470 160 80 68
CPE1 (nF) 104 0.2 1 103
α1 0.92 0.94 0.92 0.90
Ferrite/Y2O3 12 Rs (Ω) 200 300 400 450
R1 (Ω) × 102 112 49 16 9
CPE1 (nF) 10 10 0.08 0.5
α1 0.88 0.89 0.96 0.86
Ferrite/Gd2O3 16 Rs (Ω) 390 900 391 1110
R1 (Ω) × 103 500 110 10 10
CPE1 (μF) 10−3 4.20 1 100
α1 0.8 0.90 0.98 0.90
R2 (Ω) × 103 300 20 40 4
CPE2 (μF) 3 × 10−3 200 1 90
α2 1 0.80 0.88 1
Ferrite/CaO 6 Rs (Ω) 237 247 259 265
R1 (Ω) 10 19 19 5
CPE1 (μF) 20 20 20 20
α1 0.90 1 0.98 0.95
R2 (Ω) × 102 106 90 60 11
CPE2 (μF) 50 70 70 70
α2 0.87 0.94 0.98 0.90
Cole–Cole diagrams measured at selected temperatures for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C. The red thin solid line denotes the simulated diagrams and the inset is the equivalent electrical circuit using Zview program.
Figure 8
Cole–Cole diagrams measured at selected temperatures for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C. The red thin solid line denotes the simulated diagrams and the inset is the equivalent electrical circuit using Zview program.
Arrhenius plots of Co0.2Ni0.3Zn0.5Fe2O4 and its associated doped samples sintered at 1000 °C. In the X-axis labeling the symbol T is the absolute temperature expressed in Kelvin.
Figure 9
Arrhenius plots of Co0.2Ni0.3Zn0.5Fe2O4 and its associated doped samples sintered at 1000 °C. In the X-axis labeling the symbol T is the absolute temperature expressed in Kelvin.
Table 4 Selected electrical/dielectric characteristics of samples sintered at 1000 °C under a pressure of 7 tons cm−2 of the undoped ferrite, Co0.2Ni0.3Zn0.5Fe2O4, and their associated doped ones. Data of Co0.2Ni0.3Zn0.5Fe2O4 were reported by us in a previous work (Huili et al., 2015b).
Undoped ferrite Ferrite/Y2O3 Ferrite/Gd2O3 Ferrite/CaO
Ef/eV 0.27 0.91 0.75 0.39
EP/eV 0.96 1.78 1.25 1.51
TC/K 549 595 516 523
ρdc (25 °C)/Ω cm 1.72 × 109 1.57 × 1011 1.41 × 1011 5.56 × 106
fr (250 °C) × 104/Hz 2.24 200 22.4 >13 MHz
τr (250 °C)/sec 7.11 × 10−6 7.96 × 10−8 7.11 × 10−7
tan δ (123 °C & 13 MHz) 10−4 1.24 × 10−5 1.17 × 10−2 1.28
Frequency dependence of ac conductivity, σac, measured at selected temperatures for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C.
Figure 10
Frequency dependence of ac conductivity, σac, measured at selected temperatures for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C.
Frequency dependence of the real part of the electric modulus, M′, measured at selected temperatures for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C.
Figure 11
Frequency dependence of the real part of the electric modulus, M′, measured at selected temperatures for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C.
Frequency dependence of the imaginary part of the electric modulus, M″, measured at selected temperatures for the undoped ferrite, Co0.2Ni0.3Zn0.5Fe2O4, and its associated doped samples sintered at 1000 °C.
Figure 12
Frequency dependence of the imaginary part of the electric modulus, M″, measured at selected temperatures for the undoped ferrite, Co0.2Ni0.3Zn0.5Fe2O4, and its associated doped samples sintered at 1000 °C.

As can be noticed, for a given sample, as the temperature increases, fr shifts toward higher frequencies implying a decrease in the relaxation time consequence of thermal activation of charge carriers. In addition, fr was found to be higher with the doped samples. Further it depends on the nature of doping oxide. The high relaxation frequency (>13 MHz) is observed with CaO doping. It is important to note that the trend in fr correlates the ac and dc conductivities trends observed for the studied samples; the higher the relaxation frequency is the larger conductivity will be. The last conclusion allows giving another way for the explanation and prediction of the ac and dc conductivities trends.

Frequency dependence of the imaginary dielectric constant, ε“, and dielectric loss factor, tan δ, at selected temperatures are shown in Figs. 13 and 14. It is clear that all the studied samples exhibit dielectric dispersion where the both ε” and tan δ decrease rapidly with increasing frequency of the applied field. Beyond a certain critical value of frequency, the dielectric properties become almost constant. The above dielectric dispersion is a general trend observed for all ferrites. It is explained on the basis on the two polarization mechanisms (Maxwell, 1973; Wagner, 1913; Shinde et al., 2008; Sarah and Suryanarayana, 2003): (i) electrons exchange between the Fe3+ and Fe2+ ions at octahedral sites and (ii) space charge polarization due to the presence of an inhomogeneous dielectric structure. For mechanism (i) which is expected to play the dominant role, the polarization occurs through a mechanism similar to the conduction process; the exchange of electrons between the Fe3+ and Fe2+ ions at octahedral sites may lead to the local displacement of electrons in the direction of the field applied and these electrons determine the polarization. Moreover, owing to the comparatively greater polarizability of the Fe2+ ions which have a greater number of electrons than the Fe3+ ions, it can be understood that variation in dielectric constant of ferrites strongly correlated with the variations in the concentration of Fe2+ ions. At low frequencies, the polarization can follow the applied field and therefore a high dielectric constant and loss factor are recorded. With the increase in frequency, ε“ and tan δ decreased substantially and reached an almost constant value beyond certain frequency of external field (see Fig. 13). This can be explained by the fact that beyond a critical frequency the hopping frequency of the electron exchange between Fe2+ and Fe3+ cannot follow the alteration of ac electric field. Mechanism (ii) suggested as outlined before that the sample can be modelized by the M-W model consisting of a heterogeneous structure containing well conducting grains separated by highly resistive thin grain boundaries. This causes localized accumulation of charge under the applied electric field which builds up space charge polarization. Hence a high value of dielectric constant is expected at low frequency (Shinde et al., 2008; Ajmal and Maqsood, 2007). The electron reverses their direction with the increase in field reversal frequency of electric field. The chances of electron accumulation at grain boundaries decreased thereby decreasing polarization. Therefore, the dielectric features ε” and tan δ decrease with rise in frequency and reach almost constant value as is observed. For our samples the lowest value of tan δ is observed for the composition Co0.2Ni0.3Zn0.5Fe2O4/Y2O3 and it is found to be remarkably lower than that reported for the unsubstituted undoped bulk ferrite, Ni0.5Zn0.5Fe2O4 (Huili et al., 2014; Shokrollahi and Janghorban, 2007). The improved electrical/dielectric properties obtained for the Y2O3 doped ferrite can be attributed to a number of interdependent factors including the curtailing of the Fe2+ ions, the better compositional stoichiometry, the moderate sintering conditions, and fine grained pre- and post-sintered component materials. The low loss values at higher frequencies show the potential applications of these materials, especially the Y2O3 doped material in high-frequency microwave devices. As shown in Fig. 14, the CaO doped ferrite, exhibited an additional peak at about 35 kHz. The phenomenon is known as ferromagnetic resonance. It has been observed with other ferrites (Singh et al., 2011; Khan et al., 2014; Junaid et al., 2016; Iqbal et al., 2014). Ferromagnetic resonance can be observed if an ion has two equilibrium states separated by a potential barrier, and the jumping probability of both ions will be the same. The frequency that changes the position of ion is called the natural frequency of that ion (fN). When both the natural and external applied frequencies were the same, then the maximum electrical energy was transferred to the oscillating ions and then the power loss increased resulting in the resonance phenomenon. In our case, since ferromagnetic resonance was observed only with the ferrite/CaO, it can be suggested that the calcium ions in their crystal lattice occupy two energetically equal equilibrium states separated by a potential barrier.

Frequency dependence of the imaginary part of dielectric permittivity, ε″, for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C.
Figure 13
Frequency dependence of the imaginary part of dielectric permittivity, ε″, for (a) undoped ferrite, (b) ferrite/Y2O3, (c) ferrite/Gd2O3, and (d) ferrite/CaO, sintered at 1000 °C.
Frequency dependence of the loss factor, tan δ, measured at 123 °C for Co0.2Ni0.3Zn0.5Fe2O4/additives sintered at 1000 °C. The inset is a zoom-view at high frequencies.
Figure 14
Frequency dependence of the loss factor, tan δ, measured at 123 °C for Co0.2Ni0.3Zn0.5Fe2O4/additives sintered at 1000 °C. The inset is a zoom-view at high frequencies.

4

4 Conclusion

The influence of fine particles of three oxide (Y2O3, Gd2O3 or CaO) on the microstructure and the electromagnetic properties of sintered nanosized Co0.2Ni0.3Zn0.5Fe2O4 was investigated. Pure phases of both nanoadditives and ferrite were successfully produced via the polyol method. The porosity was found to decrease with the additives incorporated into the ferrite material. The lowest porosity was found with CaO (∼6% vs 18% for the undoped ferrite). dc magnetic study of the ferrite/nanoadditives showed reasonable Msat (∼70 emu g−1 at 300 K), high TC (∼ 600 K) and low magnetization decay on increase in temperature. The electrical properties of ferrites were studied at different temperatures and frequencies. Clearly all the additives affected the electrical and dielectric properties of the undoped ferrite. For the Y2O3 doped material, improved electrical characteristics (ρdc = 1.57 × 1011 Ω cm at 25 °C, tan δ = 1.24 · 10−5 at 123 °C & 13 MHz) compared with those of the undoped ferrite as well as those of similar bulk ferrites were obtained. The superior electrical/dielectric features of the material take advantages of the high quality powders of both the ferrite and the additive in terms of microstructure, chemical homogeneity, and low concentration of Fe3+/Fe2+ pair consequence of the moderate sintering conditions. The present investigation clearly demonstrated the merit of the additive Y2O3 to produce high-quality ferrite cores capable of operating at increasingly higher frequencies even in high temperature environment. Additional research work devoted to the optimization of Y2O3 doping content in the host ferrite would further improve the electromagnetic properties of the Co substituted Ni-Zn nanoferrite.

Acknowledgment

The authors are thankful to the teams of Prof. H. Boughzala (Faculty of Sciences of Tunis, University of Tunis El Manar, Tunisia), and Prof. A. Madani (Faculty of Sciences of Bizerte, University of Carthage, Tunisia) for access to the XRD and electric impedance spectroscopy equipment, respectively.

References

  1. , , , , . Structural and magnetic properties of Sr0.5Co0.5Fe2O4 nanoferrite. J. Magn. Magn. Mater.. 2014;365:83-87.
    [Google Scholar]
  2. , . Economics of ceramic magnets. Am. Ceram. Soc. Bull.. 1994;73:62-65.
    [Google Scholar]
  3. , , . Influence of zinc substitution on structural and electrical properties of Ni1−xZnxFe2O4 ferrites. Mater. Sci. Eng., B. 2007;139:164-170.
    [Google Scholar]
  4. , , , . Refinement of two-phase Zr-doped Y2O3. Mater. Sci. Forum. 1998;278:680-685.
    [Google Scholar]
  5. , , . Finite-size effects in fine particles: magnetic and transport properties. J. Phys. D Appl. Phys.. 2002;35:15-42.
    [Google Scholar]
  6. , , . Infrared absorption spectra of rare-earth oxides in the region 800 to 240 cm−1. J. Am. Ceram. Soc.. 1963;46 294-294
    [Google Scholar]
  7. Beauclair, F., Delvinquier, J.P., GROS, J.P., 1998. Transformateurs et inductances, technique de l’ingénieur.
  8. , , . Evaluation of a long-time temperature drift in a commercial Quantum Design MPMS SQUID magnetometer using Gd2O3 as a standard. J. Magn. Magn. Mater.. 2006;299:281-287.
    [Google Scholar]
  9. , , , . Choice of collimators for a crystal spectrometer for neutron diffraction. Nucl. Inst.. 1958;3:223-228.
    [Google Scholar]
  10. , . Computer Program FULLPROF. Grenoble: Laboratoire Léon Brillouin CEA-CNRS; .
  11. , . Introduction to Magnetic Materials. New York: Addison-Wesley; .
  12. , . Modern Ferrite Technology (second ed.). New York: Springer Science+Business Media; .
  13. , , , , , . Structural, magnetic and electrical properties of Co1−xZnxFe2O4 synthesized by co-precipitation method. J. Magn. Magn. Mater.. 2007;311:494-499.
    [Google Scholar]
  14. , , , , . Effect of cobalt substitution on the structure, electrical, and magnetic properties of nanocrystalline Ni0.5Zn0.5Fe2O4 prepared by the polyol process. Ceram. Int.. 2014;40:16235-16244.
    [Google Scholar]
  15. , , , . Polyol-made stoichiometric Co0.2Ni0.3Zn0.5Fe2O4 nanoparticles: synthetic optimization, structural, and microstructural studies. Int. J. Nanotechnol.. 2015;12:631-641.
    [Google Scholar]
  16. , , , , , . Effect of sintering conditions on the structural, electrical, and magnetic properties of nanosized Co0.2Ni0.3Zn0.5Fe2O4. Ceram. Int.. 2015;41:6212-6225.
    [Google Scholar]
  17. , , , , , , . High frequency dielectric properties of Eu3+ substituted Li–Mg ferrites synthesized by sol–gel auto-combustion method. J. Alloys Compd.. 2014;586:404-410.
    [Google Scholar]
  18. Johnson, D., 2005. Zview, Version 2.9c, Scribner Associates Inc, North Carolina, USA.
  19. , . Hopping losses in polarisable dielectric media. Nature. 1974;250:191-193.
    [Google Scholar]
  20. , , , , , , , , . Structural, spectral, dielectric and magnetic properties of Tb–Dy doped Li-Ni nano-ferrites synthesized via micro-emulsion route. J. Magn. Magn. Mater.. 2016;419:338-344.
    [Google Scholar]
  21. , , , , , , , , , . High frequency dielectric response and magnetic studies of Zn1−xTbxFe2O4 nanocrystalline ferrites synthesized via micro-emulsion technique. J. Magn. Magn. Mater.. 2014;360:188-192.
    [Google Scholar]
  22. , . On the dispersion of resistivity and dielectric constant of some semiconductors at audiofrequencies. Phys. Rev.. 1951;83:121-124.
    [Google Scholar]
  23. , . Monte Carlo indexing with McMaille. Powder Diffr.. 2004;19:249-254.
    [Google Scholar]
  24. Lebourgeois, R., Perriat, P., Labeyrie, M., 1992. High and Low Level Frequency Losses in NiZn and MnZn Spinel Ferrites, ICF-6 Tokyo, Japan.
  25. , , , . Phys. Chem. Glasses. 1972;13:171.
  26. , . Electricity and Magnetism (second ed.). New York: Oxford University Press; .
  27. , , . Infrared spectra and characteristic frequencies of inorganic ions. Anal. Chem.. 1952;24:1253-1294.
    [Google Scholar]
  28. , . Influence of PbO and TiO2 additives on the microstructure development and magnetic properties of Ni–Zn soft ferrites. J. King. Saud. Univ.. 2014;26:152-158.
    [Google Scholar]
  29. , , , , . Dielectric properties of polycrystalline mixed nickel–zinc ferrite. Mater. Lett.. 1999;40:39-45.
    [Google Scholar]
  30. , , . Electrical conductivity of some nickel-zinc ferrites. Phys. Stat. Sol. (A). 1976;18:647-651.
    [Google Scholar]
  31. , , , , , , . Electrical and switching properties of NiAlxFe2−xO4 ferrites synthesized by chemical method. Phys. B. 2011;406:663-668.
    [Google Scholar]
  32. , . Magnetic fluids: fabrication, magnetic properties, and organization of nanocrystals. Adv. Funct. Mater.. 2001;11:323-336.
    [Google Scholar]
  33. , , , . Electrical conductivity and dielectric behaviour of nanocrystalline NiFe2O4 spinel. J. Phys.: Condens. Matter. 2002;14:3221-3237.
    [Google Scholar]
  34. , , , , , . Microstructure and magnetic properties of microwave sintered NiCuZn ferrite for application in LTCC devices. Mater. Lett.. 2012;79:103-105.
    [Google Scholar]
  35. Rahaman, M.N., 2003. Ceramic Processing and Sintering, second ed., Materials Engineering.
  36. , . Sintering of Ceramics. CRC Press; .
  37. , , , , . Dielectric behavior of niobium doped Ni-Zn ferrites. J. Optoelectron. Adv. Mater.. 2004;6:951-954.
    [Google Scholar]
  38. , . Dependence of electrical conductivity (σ) and activation energy of lithium ferrite on sintering temperature and porosity. Crysr. Res. Technol.. 1991;26:457-464.
    [Google Scholar]
  39. , , , , . Influence of additives on the properties of a Ni–Zn ferrite with low Curie point. J. Magn. Magn. Mater.. 2000;215–216:194-196.
    [Google Scholar]
  40. , . A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr.. 1969;2:65-71.
    [Google Scholar]
  41. , , . Magnetic properties of diphasic composites of LiF5O8-BaTiO3. Indian J. Phys.. 2003;77:449-452.
    [Google Scholar]
  42. , . Determination of the size and internal structure of colloidal particles using X-rays. Nachr. Ges. Wiss. Goettingen.. 1918;2:98-100.
    [Google Scholar]
  43. , , , , , . Structural and electrical characterization of Ni-Zn ferrites. IJETCAS. 2014;7:20-25.
    [Google Scholar]
  44. , , , . DC resistivity of Ni–Zn ferrites prepared by oxalate precipitation method. Mater. Chem. Phys.. 2008;111:87-91.
    [Google Scholar]
  45. , , . Influence of additives on the magnetic properties, microstructure and densification of Mn–Zn soft ferrites. Mater. Sci. Eng., B. 2007;141:91-107.
    [Google Scholar]
  46. , , , . Dielectric relaxation, conductivity behavior and magnetic properties of Mg substituted Zn–Li ferrites. Curr. Appl. Phys.. 2011;11:783-789.
    [Google Scholar]
  47. , , . Ferrites. London: Cleaver-Hume Press; .
  48. , , . Behaviors of additives in ferrites. J. Phys. I. 1977;4:175-178.
    [Google Scholar]
  49. , , , . A domain size effect in the magnetic hysteresis of NiZn-ferrites. Appl. Phys. Lett.. 1996;69:2927-2929.
    [Google Scholar]
  50. , . J. Chem. Phys.. 1955;23:1883.
  51. , , . Electronic conductivity and transition point of magnetite (“Fe3O4”) Physica. 1941;8:979-987.
    [Google Scholar]
  52. , . Zur Theorie derunvollkommenen Dielektrika. Ann. Phys.. 1913;40:817-821.
    [Google Scholar]
  53. , , , , , . Fabrication and magnetic investigation of Y or Gd containing CoFe2O4 nanocrystalline films. J. Appl. Phys.. 1998;84:5703-5708.
    [Google Scholar]
  54. , . The crystal structure of the modification C of the sesquioxides of the rare earth metals, and of indium and thallium. Norsk. Geol. Tidsskr.. 1927;9:310-316.
    [Google Scholar]
  55. , , , , . Magnetization temperature dependence in iron nanoparticles. Phys. Rev. B. 1998;58:14167-14170.
    [Google Scholar]
  56. , , , , , , , , , . Room temperature ferromagnetism in pure Y2O3 nanoparticles. EPL. 2012;17005:1-5.
    [Google Scholar]
Show Sections