Articles
  • Preparation and characterization of hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 by the solid-state reaction route
  • Yujie Yang*, Xiansong Liu, Shuangjiu Feng, Xucai Kan, Qingrong Lv, Feng Hu, Jiangli Ni, Chaocheng Liu and Wei Wang

  • Engineering Technology Research Center of Magnetic Materials, School of Physics & Materials Science, Anhui University, Hefei 230601, P. R. China

Abstract

The microstructural, spectral, magnetic and electric properties of hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 (0.00 ¡Â x ¡Â 0.30) synthesized by the solid-state reaction route have been studied. XRD results confirmed that the hexaferrites with Nd-Zn content (x) of 0.00 ¡Â x ¡Â 0.24 were single M-type phase, and the hexaferrite with x = 0.30 exhibited the M-type phase and impurity phase. The remanence (Br) increased with x from 0.00 to 0.06, and then decreased when x ¡Ã 0.06. The intrinsic coercivity (Hcj) and magnetic induction coercivity (Hcb) decreased with x from 0.00 to 0.30. Br indicated a linear decreasing behavior with increasing temperature from 20 oC to 140 oC. Hcj raised linearly with increasing temperature from 20 oC to 140 oC. The value of aBr basically remained constant with Nd-Zn content (x). The value of aHcj increased with x from 0.00 to 0.12, and began to decrease when x ¡Ã 0.12. The electrical resitivity (¥ñ) presented a decreasing trend with x from 0.00 to 0.30.


Keywords: M-type hexaferrites, Solid-state reaction route, X-ray diffraction, Magnetic measurements

introduction

Hexagonal ferrites have received considerable attention of researchers and engineers since the discovery of the ferrite in the 1950s. Although not as powerful as the newest NdFeB or SmCo5 magnets, they still have a large market share in the market of magnetic materials, because of their perfect chemical stability, high Curie temperature, high performance-price ratio, and easy methods of production [1, 2]. In addition, M-type hexaferrites can be widely utilized as microwave devices, permanent magnetic materials, radar communication and microwave devices [3-5]. Many preparation techniques have been used to prepare the M-type hexaferrites such as hydrothermal method [6], three-step calcination method [7], co-precipitation method [8], standard solid-state reaction route [9], molten flux calcination method [10], sol-gel method [11], pulsed laser deposition method [12], and extrusion-based three-dimensional (3D) printing [13]. In the above-mentioned methods, the solid-state reaction route was used to fabricate the hexaferrites because of a few several profitable factors, for example controllable grain size, low manufacturing cost, simple technology and high productive.
In order to modify the intrinsic magnetic properties of M-type hexaferrites, the M-type hexaferrites been doped with rare-earth metals and transition metals, or the combination of these [14-36]. Singh et al. reported the rare-earth substitution (La3+, Nd3+ and Sm3+ ) doped strontium ferrite (Sr-M), and the results exhibited that the magnetization moment (Ms) and remenance (Mr) decrease with increasing rare-earth ions substitution, while the enhancement of Hc values may be due to higher magnetocrystalline anisotropy [15]. The Nd-substituted strontium hexaferrites prepared by hydrothermal synthesis have been reported by Wang et al. [17], and it is found that Nd substitution with a Nd-Sr ratio of 1/8 enhances the coercivity without causing any significant deterioration in either the saturation magnetization or the remanence. Shekhawat et al. synthesized the La-Sm substituted Sr-hexaferrite SrAl4(La0.5Sm0.5)xFe8-xO19 (0.0 ≤ x ≤ 1.5) nanomaterials by the auto combustion method and found that the remanence magnetization increases while intrinsic coercivity decreases with substitution [18]. The Zn-doped Ba hexaferrite single crystals have been synthesized, and it is found that Zn substitution significantly influences the coercivity and magnetization of Ba hexaferrite while the Curie temperature was nearly constant over the range of doping [23]. Zhang et al. reported the Nd-Co doped strontium hexaferrites Sr1-xNdxFe12-xCoxO19 (0.0 ≤ x ≤ 0.4) fabricated by sol-gel autocombustion method, and the results showed that Nd-Co substitution can improve the saturation magnetization and coercivity and reaches a maximum at x=0.2 [33]. Liu et al. prepared the Ce-Zn co-substituted M-type strontium hexaferrites by the ceramic method, and the results showed that Ms and Hc can be improved significantly [36].
In this article, in order to enhance the magnetic properties, we have selected a combination of Nd3+ and Zn2+ ions to substitute in the M-type Ba-Sr hexaferrites. The hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 (0.00 ¡Â x ¡Â 0.30) were successfully synthesized by the solid-state reaction route. Impact of Nd-Zn co-doping on the microstructural, spectral and electric properties was analyzed. Subsequently, the magnetic properties of hexaferrites with different Nd-Zn content (x) were systematically investigated under different temperatures from room temperature (20 oC) to 140 oC. The novelty of this work is doing a study on the temperature coefficient of remanence (Br) (aBr), and temperature coefficient of intrinsic coercivity (Hcj) (aHcj) for the Nd-Zn co-doped hexaferrites.

experimental details

The hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 (0.00 ≤ x ≤ 0.30) were prepared via the solid-state reaction route [37]. Barium carbonate (BaCO3), strontium carbonate (SrCO3), metallic oxides (Nd2O3, Fe2O3 and ZnO) were weighted in stoichimetric ratio. And then, the raw materials were thoroughly mixed together in water for 10 h in a ball mill in order to obtain finely mixed powder. Further, the as-mixed powder was dried, and pressed into circular pellets with Φ30 × 16 mm. Subsequently, the pellets were calcined in a muffle furnace at 1,260 oC for 2.0 h. The calcined pellets were shattered in a vibration mill, and suitable additives (CaCO3 0.8 wt%, SiO2 0.2 wt% and Al2O3 0.2 wt%) were added, then the mixture was wet-milled in a ball-mill for 16 h. This procedure guarantees a narrow particle size distribution with the mean size of around 0.75 μm. The fine milled hexaferrite slurry was pressed into circular pellets with Φ30 × 16 mm in a magnetic field of 1.2 T. In order to measure the DC electrical resitivity, the fine milled hexaferrite slurry was dried, and then pressed into circular pellets with Φ20 × 8 mm. Finally, all green pellets were sintered in a muffle furnace at 1,185 oC for 1.5 h.
The phase identification and structural characterization of synthesized samples were performed by X-ray diffraction (XRD, Rigaku Smartlab). The surface morphology of the hexaferrites was detected by a field emission scanning electron microscopy (FE-SEM, HITACHI S-4800). The magnetic properties were measured at different temperatures from room temperature (20 oC) to 140 oC using a Hysteresis graph meter (NIM-2000HF, National Institute of Metrology of China). (Resistivity testing system, Ningbo rooko FT-353) was used to measure the DC electrical resitivity (ρ) of the hexaferrites at room temperature.

results and discussion

Fig. 1 presents the XRD patterns of hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x). The characteristics peaks were compared with the M-type hexaferrite ICCD card No. 51-1879. It can be observed that the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-x     ZnxO19 with Nd-Zn content (x) ≤ 0.24 belonged to M-type hexagonal crystal structure. The presence of hematite (α-Fe2O3) (ICCD card no. 89-0599) detected for the hexaferrite with Nd-Zn content (x) = 0.30 might be due the incomplete reaction under the preparing conditions. These show that the maximum doped content of Nd-Zn co-doping can not exceed x = 0.24.
The lattice parameters c and a of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 were calculated using to the following formula [38].
 
 

 
 
In the given relation, dhkl is the interplaner spacing of the lines in XRD pattern and h, k and l are the Miller indices. The values of lattice parameters c and a for the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x) are listed in Table 1. The values of c and a decrease with increasing Nd-Zn content (x) from 0.00 to 0.30. The possible explanation for the decrease of c and a with increasing Nd-Zn content (x) can be attributed to the difference in the ionic radii (Δr) of the metal ions and the number of ionic substitutions of each species. Substitution of Sr2+ (r = 1.180 Å) by Nd3+ (r = 0.983 Å) makes a negative difference in the ionic radii of Δr = -0.197 Å. Substitution of Fe3+ (r = 0.645 Å) by Zn2+ (r = 0. 740 Å) makes a positive difference in the ionic radii of Δr = +0.095 Å. As in the case of Nd-Zn content (x) = 0.12, the lattice parameter c and a are decreased.
The unit cell volume (Vcell) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 was calculated from the following equation [38]:
 
 

 
And its values are summarized in Table 1. It is clearly seen that the unit cell volume (Vcell) decreases with Nd-Zn content (x) from 0.00 to 0.30. The decrease in the Vcell is due to the decrease of lattice parameters c and a with increasing Nd-Zn content (x) as shown in Table 1. The X-ray density (dxrd) has been calculated by using the below formula [39]:
 
 

 
 
where Z is the number of molecular per unit cell, M is the molecular weight, NA is Avogadro’s number and Vcell is unit cell volume. And the values dxrd are listed in Table 1. It noticed that the value of dX-ray enhances from 5.194 g/cm3 at x = 0.00 to 5.321 g/cm3 at x = 0.30. The enhancement of dxrd with increasing Nd-Zn content (x) can be attributed to the larger molecular weight and decreasing trend of unit cell volume (Vcell) for the Nd-Zn co-doped hexaferrites displayed in Table 1.
Fig. 2 shows the FE-SEM images of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with x = 0.00, x = 0.12, and x = 0.24. This indicates that the grains in the hexaferrites are homogeneous distribution with hexagonal plate like shape. The average grain size increases from 1.8 μm at x = 0.00 to 3.2 μm at x = 0.24.
Fig. 3 represents the room temperature demagnetizing curves of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x). As revealed by the demagnetizing curves, the magnetic properties of the hexaferrites are greatly influenced by the substitution of Sr2+ ion by Nd3+ ion and Fe3+ ion by Zn2+ ion. The remanence (Br), magnetic induction coercivity (Hcb), and intrinsic coercivity (Hcj) at the room temperature are derived from the demagnetizing curves of the hexaferrites samples with different Nd-Zn content (x).
The remanence (Br) of the hexaferrites Ba0.40Sr0.60-x                  NdxFe12.00-xZnxO19 as a function of Nd-Zn content (x) is plotted in Fig. 4(a). It is clearer that a suitable amount of Nd-Zn substitution can increase the remanence (Br) of the hexaferrites. In Fig. 4(a), with increasing Nd-Zn content (x), Br first increases and reaches to the maximum value of 386.0 mT at Nd-Zn content (x) = 0.06, and then decreases when Nd-Zn content (x) ≥ 0.06. The changing trend of remanence (Br) can be explained on the basis of the occupation of the substituted ions and the impurity phase. M-type hexaferrites belong to P63/mmc space group, and have five crystallographic sublattices, such as three kinds of octahedral sites (2a, 12k, 4f2), one tetrahedral site (4f1) and one bipyramidal site (2b). The 2a, 12k and 2b sites have upward spin direction, and the 4f1 and 4f2 sites have downward spin direction [37, 40]. Herme et al. [41] have reported that in the M-type hexaferrites, Nd3+ ions substitute the Sr2+ sites in the vicinity of 4f2 sites via Mössbauer analysis. Lee et al. [42] have reported that Zn2+ ions prefer to occupy the 4f1 and 2b sites of tetrahedral sites in the M-type hexagonal structure by Mössbauer spectra. Zn2+ ion is nonmagnetic. And the magnetic moment of Fe3+ ion is 5 μB. Therefore, the increase of remanence (Br) for the hexaferrites with increasing Nd-Zn content (x) from 0.00 to 0.06 is probably due to the below factor. When Nd-Zn content (x) ≤ 0.06, the number of Zn2+ ions entering 4f1 sites (spin down) is more than that of Zn2+ ions entering 2b sites (spin up). This in turn decreases the negative magnetic moment of Fe3+ ions. Consequently, according to equation (3), the whole magnetic moment is enhanced. This leads to the increase of the remanence (Br). When Nd-Zn content (x) ≥ 0.06, the decrease of remanence (Br) for the hexaferrites may be attributed to the below three reasons. Firstly, at Nd-Zn content (x) ≥ 0.06, Zn2+ ions substitute Fe3+ ions in 2b sites having spin-up. According to equation (3), this in turn decreases the whole magnetic moment. This results in the decrease of the remanence (Br). Secondly, Zn2+ (0 μB) ions substituting magnetic Fe3+ ions (5 μB) weaken the super-exchange interaction between metallic ions in the M-type hexaferrites, and then the remanence (Br) is decreased. Thirdly, when Nd-Zn content (x) = 0.30, as seen from Fig. 1, the impurity phase (hematite: α-Fe2O3) leads to the decrease of remanence (Br).
Fig. 4(b) shows the intrinsic coercivity (Hcj) and magnetic induction coercivity (Hcb) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 as a function of Nd-Zn content (x). As shown in Fig. 4(b), the values of Hcj and Hcb show a decreasing trend with increasing Nd-Zn content (x), and decrease from 225.0 and 215.9 kA/m at x = 0.00 to 126.2 and 120.7 kA/m at x = 0.30, respectively. This proposes that Nd-Zn co-substitution can modify the coercivity of M-type Ba-Sr hexaferrites. Yang et al. [43] have reported that in the M-type hexaferrites, the ions at octahedral 4f2 site and bipyramidal 2b site are known to be the main contributors to the magnetocrystalline anisotropy. According to the Stroner-Wohlfarth theory, the intrinsic coercivity (Hcj) of the M-type hexaferrites which results from coherent rotation of magnetization can be calculated by the following equation [44]:
 
 

 
where C is dimensionless constant of material, K1 is the first anisotropy constant, and N is the grain demagnetization factor. Therefore, according to the equation (4), the decrease of Hcj with increasing Nd-Zn content (x) from 0.00 to 0.30 could be due to the following two reasons. Firstly, as seen from Fig. 2, the average grain size of M-type hexaferrite increases with increasing Nd-Zn content (x) from 0.00 to 0.30. Thus, the value of N increases with increasing Nd-Zn content (x). According to equation (4), the value of Hcj is decreased. Secondly, the results of Mössbauer spectra showed that Zn2+ ions prefer to occupy the 4f1 and 2b sites of tetrahedral sites in the M-type hexagonal structure [42]. The decrease of Hcj with increasing Nd-Zn content (x) may be related to the reduction of magnetocrystalline anisotropy field as a result of Zn2+ substitution for Fe3+ ions at 2b sites.
The temperature dependent demagnetizing curves of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x) are presented in Fig. 5. The changing trend of the demagnetizing curves measured at different temperatures is in agreement with that reported by T.D.K. Corporation [45] and A. Goldman et al. [46]. The values of the remanence (Br), and intrinsic coercivity (Hcj) at different temperatures are calculated from the demagnetizing curves for the hexaferrites with different Nd-Zn content (x). Fig. 6(a) presents the temperature dependent remanence (Br) between 20 oC and 140 oC for the hexaferrites with different Nd-Zn content (x). As seen from Fig. 6(a), for the hexaferrites with different Nd-Zn content (x), the values of remanence (Br) have a linear decreasing behavior with increasing temperature from 20 oC to 140 oC. This is in agreement with the results reported by Zhou et al. [47]. The variations of the intrinsic coercivity (Hcj) between 20 oC and 140 oC for the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x) are shown in Fig. 6(b). It is observed that for the hexaferrites with different Nd-Zn content (x), the values of intrinsic coercivity (Hcj) raise linearly with increasing temperature from 20 oC to 140 oC. This agrees with the changing trend reported by W. Zhou [48] and T.D.K. Corporation [49].
When the ambient temperature ¡Â 400 oC, the remanence (Br) of hexaferrites BaFe12O19 deareases linearly with increasing temperature [50]. The temperature coefficient of remanence (Br) can be approximated as a constant in a certain temperature range. Therefore, the average temperature coefficient of Br (aBr) of hexaferrites can be defined as the below relation [48]:
 

 
where Br(20) is the value of remanence (Br) measured at 20 oC, Br(t) is the value of remanence (Br) measured at temperature of t oC. Fig. 7(a) represents the variation of aBr of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 as a function of Nd-Zn content (x). It is clear from Fig. 7(a) that the value of aBr basically remains constant around -0.175 %/oC. The value of aBr is in agreement with that reported by T.D.K. Corporation [45]. This indicates that the Nd-Zn substitution have not big impact on the average temperature coefficient of remanence (Br) (aBr) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19.
As shown in Fig. 6(b), the values of intrinsic coercivity (Hcj) have a linear behavior with temperature from 20 oC to 140 oC. The temperature coefficient of intrinsic coercivity (Hcj) is also approximated as a constant in a certain temperature range [49]. Thus, the average temperature coefficient of Hcj (aHcj) can be calculated using the following equation [43]:
 

 
where Hcj(20) is the value of intrinsic coercivity (Hcj) measured at 20 oC, Hcj(t) is the value of intrinsic coercivity (Hcj) measured at temperature of t oC. The influence of Nd-Zn content (x) on the average temperature coefficient of intrinsic coercivity (Hcj) (aHcj) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 is plotted in Fig. 7(b). It is noted that the value of aHcj firstly increases from 0.417 %/oC at x=0.00 to 0.573 %/oC at x = 0.12, and then decreases when Nd-Zn content (x) ¡Ã 0.12. This shows that Nd-Zn content (x) can significantly affect the values of aHcj. The values of aHcj are larger than that reported by T.D.K. Corporation [49], which should the values of intrinsic coercivity (Hcj) in this study are lower than that in TDK Products Catalog [49].
Fig. 8 illustrates the impact of Nd-Zn content (x) on the DC electrical resistivity (ρ) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19. It is clear that the electrical resistivity (ρ) is strongly affected by Nd-Zn content (x). It is observed that the value of electrical resitivity (ρ) decreases from 679.9 × 107 Ω-Cm at x = 0.00 to 0.13 × 107 Ω-Cm at x = 0.30. The mechanism of conductivity in the hexaferrites is attributed to the hopping of electrons between Fe3+ and Fe2+ ions at the octahedral sites [50]. Van Diepen et al. have reported that the substitution of La3+ for Ba2+ or Sr2+ in M-type ferrites is associated with a valence change of Fe3+ to Fe2+ at 2a or 4f2 site [51]. Thus, as Nd3+ ions substitute Sr2+ ions and Zn2+ ions substitute Fe3+ ions, some Fe3+ ions will change into Fe2+ ions. This increases the number of Fe2+ ions which leads to the increase of the hopping probability between the Fe3+ and Fe2+ ions. Thus, the above reasons result in the decrease of electrical resitivity (ρ).

Fig. 1

XRD patterns of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x).

Fig. 2

FE-SEM images of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with (a) x=0.00, (b) x=0.12, and (c) x=0.24.

Fig. 3

Room temperature demagnetizing curves of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with (a) x = 0.00, (b) x = 0.06, (c) x = 0.12, (d) x = 0.18, (e) x = 0.24, and (f) x = 0.30.

Fig. 4

(a) Remanence (Br), and (b) Intrinsic coercivity (Hcj) and magnetic induction coercivity (Hcb) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 as a function of Nd-Zn content (x).

Fig. 5

Measuring temperature dependent demagnetizing curves of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with (a) x=0.00, (b) x=0.06, (c) x=0.12, (d) x=0.18, (e) x=0.24, and (f) x=0.30.

Fig. 6

(a) The temperature dependent remanence (Br), and (b) The temperature dependent intrinsic coercivity (Hcj) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x) from x=0.00 to x=0.30, obtained from demagnetizing curves.

Fig. 7

(a) The average temperature coefficient of remanence (Br) (aBr), and (b) The average temperature coefficient of intrinsic coercivity (Hcj) (aHcj) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 as a function of Nd-Zn content (x).

Fig. 8

DC electrical resistivity (ρ) of the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 as a function of Nd-Zn content (x).

Table 1

The values of molecular weight, lattice parameters (c and a), unit cell volume (Vcell), X-ray density (dxrd), and bulk density (dbulk) for the hexaferrites Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 with different Nd-Zn content (x).

conclusion

The solid-state reaction route was used to synthesize the Nd-Zn substituted M-type hexaferrites with nominal compositions Ba0.40Sr0.60-xNdxFe12.00-xZnxO19 (0.00 ¡Â x ¡Â 0.30). The X-ray diffraction patterns show that the hexaferrites with Nd-Zn content (x) of 0.00 ¡Â x ¡Â 0.24 were single M-type phase, while the hexaferrites with Nd-Zn content (x) = 0.30 exhibited the M-type phase and impurity phases. FE-SEM images proposed that all the particles were regular hexagonal platelet-like shape and the average particle size increased with increasing Nd-Zn content (x). Br increased with Nd-Zn content (x) from 0.00 to 0.06, and then started to decrease when Nd-Zn content (x) ¡Ã 0.06. Hcj and Hcb decreased with Nd-Zn content (x) from 0.00 to 0.30. Br indicated a linear decreasing behavior with increasing temperature from 20 oC to 140 oC. Hcj raise linearly with increasing temperature from 20 oC to 140 oC. The value of aBr basically remained constant with Nd-Zn content (x). The value of aHcj firstly increased with x from 0.00 to 0.12, and then decreased when x ¡Ã 0.12. The electrical resitivity (¥ñ) presented a decreasing trend with x from 0.00 to 0.30.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Nos. 51872004, 51802002), Education Department of Anhui Province (Nos. KJ2013B293, KJ2018A0039), Key Program of the Science and Technology of Anhui Province (Grant No. S201904a09020074).

References
  • 1. S.V. Trukhanov, A.V. Trukhanov, V.A. Turchenko, An.V. Trukhanov, E.L. Trukhanova, D.I. Tishkevich, V.M. Ivanov, T.I. Zubar, M. Salem, V.G. Kostishyn, L.V. Panina, D.A. Vinnik, S.A. Gudkova, Ceram. Int. 44 (2018) 290-300.
  •  
  • 2. L.A. Trusov, E.A. Gorbachev, V.A. Lebedev, A.E. Sleptsova, I.V. Roslyakov, E.S. Kozlyakova, A.V. Vasiliev, R.E. Dinnebier, M. Jansen, P.E. Kazin, Chem. Commun. 54 (2018) 479-482.
  •  
  • 3. T.T. Loan, T.T.V. NGA, N.P. Duong, S. Soontarnon, T.D. Hien, J. Electron. Mater. 46 (2017) 3396-3405.
  •  
  • 4. I. Ali, M.U. Islam, M.S. Awan, M. Ahmad, J. Electron. Mater. 43 (2014) 512-521.
  •  
  • 5. R. Valenzuela, in “Magnetic ceramics” (Cambridge University Press, 1994) p.191.
  •  
  • 6. D. Makovec, D. Primc, S. Šturm, A. Kodre, D. Hanžel, M. Drofenik, J. Solid State Chem. 196 (2012) 63-71.
  •  
  • 7. S. Mahadevan, S.B. Narang, P. Sharma, Ceram. Int. 45 (2019) 9000-9006.
  •  
  • 8. N. Tran, H.S. Kim, T.L. Phan, D.S. Yang, B.W. Lee, Ceram. Int. 44 (2018) 12132-12136.
  •  
  • 9. J. Dho, E.K. Lee, J.Y. Park, N.H. Hur, J. Magn. Magn. Mater. 285 (2005) 164-168.
  •  
  • 10. C. Serletis, G. Litsardakis, E. Pavidou, K.G. Efthimiaadis, Physica B. 525 (2017) 78-83.
  •  
  • 11. Z.H. Hua, S.Z. Li, Z.D. Han, D.H. Wang, M. Lu, W. Zhong, B.X. Gu, Y.W. Du, Mater. Sci. Eng. A 448 (2007) 326-329.
  •  
  • 12. J.E. Beevers, C.L. Love, V.K. Lazarov, S.A. Cavill, H. Izadkhah, C. Vittoria, R. Fan, G. van der Laan, S.S. Dhesi, App. Phys. Lett. 112 (2016) 082401.
  •  
  • 13. X.X. Wei, Y.H. Liu, D.J. Zhao, X.W. Mao, W.Y. Jiang, S.Z. Sam Ge, J. Magn. Magn. Mater. 493 (2020) 165664.
  •  
  • 14. J.F. Wang, C.B. Ponton, I.R. Harris, J. Alloys Compd. 403 (2005) 104-109.
  •  
  • 15. A. Singh, S.B. Narang, K. Singh, O.P. Pandey, R.K. Kotnala, J. Ceram. Process. Res. 11[2] (2010) 241–249.
  •  
  • 16. A. Thakur, R.R. Singh, P.B. Barman, J. Magn. Magn. Mater. 326 (2013) 35-40.
  •  
  • 17. J.F. Wang, C.B. Ponton, I.R. Harris, IEEE Trans. Magn. 38 (2002) 2928-2930.
  •  
  • 18. D. Shekhawat, A.K. Singh, P.K. Roy, J. Mol. Struct. 1179 (2019) 787-794.
  •  
  • 19. Y.J. Yang, J.X. Shao, F.H. Wang, D.H. Huang, J. Ceram. Process. Res. 18[5] (2017) 394–398.
  •  
  • 20. J.F. Wang, C.B. Ponton, I.R. Harris, J. Magn. Magn. Mater. 298 (2006) 122-131.
  •  
  • 21. P.A. Pawar, S.S. Desai, Q.Y. Tamboli, S.E. Shirsath, S.M. Patange, J. Magn. Magn. Mater. 378 (2015) 59-63.
  •  
  • 22. G. Litsardakis, I. Manolakis, C. Serletis, K.G. Efthimiadis, J. Magn. Magn. Mater. 316 (2007) 170-173.
  •  
  • 23. D.A. Vinnik, A.S. Semisalova, L.S. Mashkovtseva, A.K. Yakushechkina, S. Nemrava, S.A. Gudkova, D.A. Zherebtsov, N.S. Perov, L.I. Isaenko, R. Niewa, Mater. Chem. Phys. 163 (2015) 416-420.
  •  
  • 24. X.-S. Liu, L. Fetnandez-Garcia, F. Hu, D.-R. Zhu, M. Suárez, J.L. Menéndez, Mater. Chem. Phys. 133 (2012) 961-964.
  •  
  • 25. M. Ghimire, S. Yoon, L. Wang, D. Neupane, J. Alam, S.R. Mishra, J. Magn. Magn. Mater. 454 (2018) 110-120.
  •  
  • 26. D.A. Vinnik, A.Yu. Tarasova, D.A. Zherebtsov, L.S. Mashkovtseva, S.A. Gudkova, S. Nemrava, A.K. Yakushechkina, A.S. Semisalova, L.I. Isaenko, R. Niewa, Ceram. Int. 41 (2015) 9172-9176.
  •  
  • 27. J. Qiu, Y. Wang, M. Gu, Mater. Lett. 60 (2006) 2728-2732.
  •  
  • 28. I.A. Auwal, H. Güngüneş, A. Baykal, S. Güner, S.E. Shirsath, M. Sertkol, Ceram. Int. 42 (2016) 8627-8635.
  •  
  • 29. A. Shayan, M. Abdellahi, F. Shahmohammadian, S. Jabbarzare, A. Khandan, H. Ghayour, J. Alloys Compd. 708 (2017) 538-546.
  •  
  • 30. H.M. Khan, M.U. Islam, Y.B. Xu, M.N. Ashiq, I. Ali, M.A. Iqbal, M. Ishaque, Ceram. Int. 40 (2014) 6487-6493.
  •  
  • 31. T.-Y. Hwang, J. Lee, H.-R. Lim, S.J. Jeong, G.-H. An, J. Kim, Y.-H. Choa, Ceram. Int. 43 (2017) 3879-3884.
  •  
  • 32. L. Peng, L.Z. Li, X.X. Zhong, Y.B. Hu, S.M. Chen, J. Magn. Magn. Mater. 428 (2017) 73-77.
  •  
  • 33. Z.Y. Zhang, X.X. Liu, X.J. Wang, Y.P. Wu, R. Li, J. Alloys Compd. 525 (2012) 114-119.
  •  
  • 34. H.M. Khan, M.U. Islam, Y.B. Xu, M. Asif Iqbal, I. Ali, M. Ishaque, M.A. Khan, J. Sol-Gel. Sci. Technol. 75 (2015) 305-312.
  •  
  • 35. Z. Lalegani, A. Nemati, J. Mater. Sci.: Mater. Electron. 26 (2015) 2134-2144.
  •  
  • 36. C.C. Liu, X.C. Kan, F. Hu, X.S. Liu, S.J. Feng, J.Y. Hu, W. Wang, K.M. Ur Rehman, M. Shezad, C. Zhang, H.H. Li, S.Q. Zhou, Q.Y. Wu, J. Alloy. Compd. 785 (2019) 452-459.
  •  
  • 37. Y.J. Yang, D.H. Huang, J.X. Shao, F.H. Wang, Chin. J. Phys. 57 (2019) 250-260.
  •  
  • 38. A. Majeed, M.A. Khan, F. ur Raheem, A. Hussain, F. Iqbal, G. Murtaza, M.N. Akhtar, I. Shakir, M.F. Warsi, J. Magn. Magn. Mater. 408 (2016) 147-154.
  •  
  • 39. V.N. Dhage, M.L. Mane, M.K. Babrekar, C.M. Kale, K.M. Jadhav, J. Alloys Compd. 509 (2011) 4394-4398.
  •  
  • 40. Y.J. Yang, X.S. Liu, D.L. Jin, Y.Q. Ma, Mater. Res. Bull. 59 (2014) 37-41.
  •  
  • 41. C. Herme, S.E. Jacobo, P.G. Bercoff, B. Arcondo, Hyperfine Interact. 195 (2010) 205-212.
  •  
  • 42. S.W. Lee, S.Y. An, In-Bo Shim, C.S. Kim, J. Magn. Magn. Mater. 290-291 (2005) 231-233.
  •  
  • 43. Z. Yang, C.S. Wang, X.H. Li, H.X. Zeng, Mater. Sci. Eng. B 90 (2002) 142-145.
  •  
  • 44. B.K. Rai, S.R. Mishra, V.V. Nguyen, J.P. Liu, J. Alloys Compd. 550 (2013) 198-203.
  •  
  • 45. TDK Electronics, in “TDK Products Catalog: Ferrite Magnets” (TDK Electronics, 2011) p.3.
  •  
  • 46. A. Goldman, in “Modern Ferrite Technology” (Springer, 2006) p.235.
  •  
  • 47. Z.G. Zhou, in “Ferrite Magnetic Materials” (Science Press, 1981) p.646.
  •  
  • 48. W.Y. Zhou, in “Design Technology of Permanent Magnet Ferrite and Magnetic Fluid” (University of Electronic Science and Technology Press, 1991) p.34.
  •  
  • 49. TDK Electronics, in “TDK Products Catalog: Ferrite Magnets” (TDK Electronics, 2014) p.4.
  •  
  • 50. F. Aen, M.F. Wasiq, M.U. Rana, H.M. Khan, H.A. Khan, Ceram. Int. 42 (2016) 16077-16083.
  •  
  • 51. A.M. Van Diepen, F.K. Lotgering, J. Phys. Chem. Solids 35 (1974) 1641-1643.
  •  

This Article

  • 2020; 21(4): 416-424

    Published on Aug 30, 2020

  • 10.36410/jcpr.2020.21.4.416
  • Received on Dec 28, 2019
  • Revised on Apr 21, 2020
  • Accepted on May 4, 2020

Correspondence to

  • Yujie Yang
  • Engineering Technology Research Center of Magnetic Materials, School of Physics & Materials Science, Anhui University, Hefei 230601, P. R. China
    Tel : +86 551 63861257
    Fax: +86 831 63861257

  • E-mail: loyalty-yyj@163.com