Co-Doped Rare-Earth (La, Pr) and Co-Al Substituted M-Type Strontium Hexaferrite: Structural, Magnetic, and Mossbauer Spectroscopy Study

Abstract

The present study investigates the influence of La3+ and Pr3+ doping on the structural, magnetic properties, and hyperfine fields of Sr0.7RE0.3Fe12-2x CoxAlxO19, (RE: La3+ and Pr3+, x = 0.0 - 0.8) hexaferrite compounds prepared via auto-combustion technique. The XRD analysis shows a linear decrease in a and c lattice and unit cell volume contraction with the content x. The room temperature magnetic study shows that for the Pr3+ doped Sr0.7Pr0.3Fe12-2x CoxAlxO19 (Pr3+-SrM), the magnetization value monotonically decreases while for La3+ doped Sr0.7La0.3Fe12-2xCoxAlxO19 (La3+-SrM) magnetization value shows a noticeable increase in magnetization value with x. The coercivity of the Pr3+-SrM compound was observed to decrease while that of the La3+-SrM compound showed a marked 40% increase at x = 0.2 (~5829 Oe) in comparison to undoped SrFe12O19 (~3918 Oe). A difference in Curie temperature was also observed, with Tc ~ 525°C at x = 0.4 for Pr3+-SrM and Tc = 505°C for x = 0.4 for La3+-SrM compound. The observed differences in magnetic properties have been explained on the basis of the site occupancy of Co2+ and Al3+ in the presence of rare-earth ions. The presence of non-magnetic rare-earth ion, La3+, improved saturation magnetization, and coercivity and deemed suitable replacement for Sr2+. The hyperfine parameters namely quadrupole shift showed a decrease with the La3+ or Pr3+ doping independent of (Co2+-Al3+) ions doping. Overall, the Mossbauer analysis suggests that the (Co2+-Al3+) impurities prefer occupancy at 2a site.

Share and Cite:

Ghimire, M. , Kunwar, D. , Dahal, J. , Neupane, D. , Yoon, S. and Mishra, S. (2020) Co-Doped Rare-Earth (La, Pr) and Co-Al Substituted M-Type Strontium Hexaferrite: Structural, Magnetic, and Mossbauer Spectroscopy Study. Materials Sciences and Applications, 11, 474-493. doi: 10.4236/msa.2020.117033.

1. Introduction

The M-type hexaferrite, Strontium hexaferrite, is an excellent candidate for technological applications because of its high uniaxial magneto-crystalline anisotropy, large magnetization, high permeability, low conductive losses, excellent chemical stability, corrosion resistance and excellent high-frequency response [1] [2] [3]. Strontium hexaferrite has been widely used as materials for industrial applications, such as in microwave devices, small motors, electromagnetic wave absorber, and ferroxdures [4] [5]. Besides this, the popularity of strontium hexaferrite is also due to its economic success, which is its low price per unit available magnetic energy and its wide availability. So far, efforts have been made to further improve electrical, dielectric and magnetic properties of strontium hexaferrite by means of doping, heat treatment, ion substitution, and processing conditions [6] [7] [8] [9] [10].

In M-type hexaferrite, the structure is comprised of 64 ions per hexagonal unit cell on 11 distinct basis sites. The 24 Fe3+ iron atoms of a unit cell occupy five different interstitial sites: three octahedral sites (2a, 12k, and 4f2), one tetrahedral site (4f1) and one trigonal bi-pyramidal site (2b). The coupling of these sites by superexchange interaction via O− 2 gives rise to a ferrimagnetic structure. The sites 12k, 2a, and 2b are sites with spin up while sites 4f2 and 4f1 are sites with the downs-spin [10]. This provides an ample opportunity to tune the magnetic properties of M-type hexaferrite compound by carefully engineering site occupancy in favor of increasing net magnetization of the compound. In view of this, attempts are made with either partial substitution of Sr2+ or Fe3+ sites. For example, substitution with non-magnetic ions such as Al3+ [11] [12], Zn3+ [13], Ga3+ [14] and Cd3+ [15] [16] and magnetic ions such as Co2+ [17] and Cr3+ [18] at Fe sites, or with the partial substitution of Sr2+ site by RE3+ such as La3+ [19] [20], Nd3+ [21], Sm3+ [22] [23] and Pr3+ [24] ions, and substitution of Sr2+/Fe3+ together with Pr-Zn [25], La-Cu [26], and La-Zn [27]. The majority of these studies have attributed changes in the magnetic behavior of doped SrM either to the site-occupancy, which perturbs exchange interaction between Fe3+-O2Fe3+, anisotropy changes occurring at 2b sites due to perturbation in electric field gradient or extrinsic features such as particle size.

Doping of non-magnetic Al3+ in SrFe12xAlxO19 has been reported to bring a considerable enhancement in the coercivity along with a reduction in saturation magnetization [11] [28] [29]. Thus with the Al3+ substitution, in spite of a large increase in the coercivity, the compound attains abysmal magnetization value. A strategy could be designed to keep both coercivity and magnetization value high. In the present work, an attempt is made to co-doped Co2+-Al3+ in the SrFe12O19 compound. Magnetic ion Co2+ (~3.7 μB) is chosen to maintain high magnetization value, while Al3+ doping is expected to maintain high coercivity value in SrFe12O19. Furthermore, Lechevallier et al. reported that the only light rare-earth enters the structure, which a solubility that is related to the shape of the charge distribution of the 4f electrons to its surroundings in the crystal structure. Thus, only those rare-earth ions have easy accommodation at the Sr2+ site, whose surroundings favor an oblate electronic distribution. This makes Pr and Nd (with oblate 4f charge distribution, negative Stevens constant) ideal candidate to occupy the Sr2+ site [25]. Furthermore, Co2+ ions were reported to modify the surroundings of the Sr2+ site, improving the introduction of rare-earth ios with oblate electronic distributions, in fact, the presence of Co2+ increases the solubility of Pr3+ [25]. Also, RE3+-Co2+-Al3+, co-substitution naturally maintains the charge neutrality condition in the compound. With the substitution of RE3+ for the Sr2+, the extra positive charge is compensated by the substitution of Co2+ at the Fe3+ site. In the absence of Co2+, one of the Fe3+ has to convert to Fe2+, which has a deleterious effect on the magnetization. With this strategy in mind, the present study investigates, the influence of the 1) rare-earth substitution, Pr3+, and La3+, at the Sr2+ site and 2) Co2+-Al3+ substitution at Fe3+ sites on the overall magnetic properties of the Sr0.7RE0.3Fe12−2xCoxAlxO19 compound.

2. Experimental

The series of Sr0.7RE0.3Fe122xCoxAlxO19 (RE: La3+ and Pr3+, x = 0, 0.2, 0.4, 0.6 and 0.8) compounds were prepared by auto-combustion method [11]. Nitrate salts of Fe(NO3)3∙9H2O (Sigma Aldrich, 99%), Sr(NO3)2 (Sigma Aldrich, 99%), Al(NO3)3∙9H2O (Sigma Aldrich, 99%), RE(NO3)3∙9H2O (Sigma Aldrich, 99%) and C6H8O7 (Sigma Aldrich, 99%) were selected for the synthesis Sr0.7RE0.3Fe122x CoxAlxO19 compounds. The stoichiometric amount of nitrates salts and citric acid were mixed in 25 ml of deionized water under constant stirring on a hot plate at 80˚C for about 30 minutes. Aqueous ammonia was added dropwise in the homogeneous mixture until the pH value of the solution reached 6.5. The intermediate precipitate was heated at 100˚C on a hot plate under stirring until the viscous brown gel is formed. The temperature is suddenly increased to 300˚C to turn it into magnetic powder. Powders were calcined in air at 1000˚C for twelve hours. The precursor used in the preparation of 1 g of sample is listed in Table 1.

The crystal structure and phase purity of the obtained samples were identified by analyzing x-ray diffraction (XRD) pattern collected via Bruker D8 Advance X-ray diffractometer using Cu Kα radiation source (λ ~ 0.154056 nm). Vibrating sample magnetometer (VSM) was used to collect hysteresis loops of the samples at room temperature in the maximum field of 1.2 T. Curie temperature, Tc, were measured using Thermogravimetric analyzer, TGA, (DuPont 910) equipped with a permanent magnet. Room temperature 57Fe Mossbauer spectroscopy was

Table 1. Stoichiometry weight of of SrFe12O19, Sr0.7La0.3Fe12-2xCoxAlxO19 and Sr0.7Pr0.3Fe12-2xCoxAlxO19.

employed to derive hyperfine parameters. The Mossbauer spectrometer (SEE Co. Minneapolis, MN USA) was calibrated against α-Fe foil. The Mossbauer spectra were analyzed using WMOSS software (SEE Co. Minneapolis, MN USA).

3. Results and Discussions

The room temperature XRD pattern of Sr0.7RE0.3Fe122xCoxAlxO19 (RE: La3+ and Pr3+, x = 0.0, 0.2, 0.4, 0.6, and 0.8) is shown in Figure 1. XRD patterns of the substituted compounds indicated the presence of a magnetoplumbite structure (ICCD 080-1198) corresponding to the hexagonal P63/mmc symmetry phase group without any secondary phase. The lattice parameters of the compound were calculated from their inter-planar spacing dhkl corresponding to the major peaks (107), (114), and (008) using equation [30]:

d ( h k l ) = ( 4 ( h 2 + h k + k 2 ) 3 a 2 + l 2 c 2 ) 1 2 (1)

where h, k, and l are Miller indices. The calculated lattice parameters a and c of the hexaferrites are shown in Table 2. The plot for a and c as a function of doping content is shown in Figure 2. A linear decrease in the lattice parameters a and c with the doping content was observed. The linear decrease in the lattice parameter may occur due to the substitution of Al3+ (ionic radii, r ~ 0.51 Å) and

(a) (b)

Figure 1. X-ray diffraction pattern of (a) Sr0.7La0.3Fe122xCoxAlxO19 and (b) Sr0.7Pr0.3Fe122xCoxAlxO19 obtained using Cu Ka radiation.

Figure 2. Lattice parameters of Sr0.7La0.3Fe122xCoxAlxO19 and Sr0.7Pr0.3Fe122xCoxAlxO19 as a function of content, x, derived from X-ray diffraction pattern.

Table 2. Lattice parameters of SrFe12O19, Sr0.7La0.3Fe122xCoxAlxO19 and Sr0.7Pr0.3Fe122xCoxAlxO19 as obtained from XRD pattern using Equation (1) and crystalline size using Equation (2).

Co2+ (r ~ 0.72 Å) for Fe3+ (r ~ 0.64 Å) and smaller La3+ (r ~ 1.172 Å) and Pr3+ (r ~ 1.13 Å) for the Sr2+ (r ~ 1.32 Å) ions [31]. According to Wagner [32], an examination of c/a parameter ratio may be used to quantify the structure type, as the M-type (magnetoplumbite) structure can be assumed if the ratio is observed to be in the range 3.917 and 3.963. As per Table 2, the c/a ratios of as-prepared samples are in the range of 3.908 to 3.919, assuring that the as-prepared samples have maintained the M-type structure.

The crystallite size of as-synthesized particles was calculated using Scherrer’s equation [33]:

D ( h k l ) = k λ / ( β cos ϑ ) (2)

where k denotes the Scherrer constant (k = 0.9), λ is the wavelength of x-ray source (λ = 0.154056 nm), β is the full-width-half-maximum of a diffraction peak and θ is the diffraction angle. As listed in Table 2, the crystallite size of as-synthesized samples is in the range of 54 to 77 nm. In the comparison of the pure SrFe12O19 sample with a crystallite size of 80.7 nm, the doped compounds show reduced crystallite size. Grain refinement is usually observed in rare-earth and doped oxide compounds [34]. This grain refinement is reported to result from the 1) increased microstrain and defect density with substitution content and 2) diffusion of substituent element to the grain boundaries, the migration that restrains the grain growth by lowering down the grain growth mobility. If the retarding force generated is more than the driving force for the grain growth due to dopants, the movement of the grain boundary is impeded [35].

The room temperature hysteresis loops, M vs. H, of Sr0.7RE0.3Fe122xCoxAlxO19 are shown in Figure 3. The magnetic parameters viz. saturation magnetization, Ms, remanence, Mr, and coercivity, Hc were extracted from the hysteresis loops and are listed in Table 3. The M vs. H curves show hard ferrites behaviors with high coercivity. The magnetic properties have been changed significantly upon RE3+ and Co2+-Al3+ substitution in Sr0.7RE0.3Fe122xCo-AlxO19 as compared to undoped ferrite SrFe12O19. The Ms value of Pr3+-SrM is observed to have a maximum value of 86 emu/g for x = 0.2 but decreases with Co2+-Al3+ content and achieves a minimum value of 65 emu/g for x = 0.8. However, in the case of La3+-SrM, the Ms increases with x and attains a maximum value of 87 emu/g at x = 0.8. As compared to pure SrFe12O19, La3+-SrM displays a noticeable 20% enhancement while Pr3+-SrM shows a 10% reduction in the Ms value. The variation of Mr and Ms as a function of x content is shown in Figure 4. The value of Mr follows a trend similar to that of Ms. As explained in the Mossbauer section, for series of compounds shows the marked preference of Co2+-Al3+ for the 2a site, a site with the spin-up moment. Despite the substitution of Co2+-Al3+ at the 2a site with the up-spin moment, the magnetization of La3+-SrM displayed a marked increase in the magnetization value with the substitution. A possible explanation for this observation is that the substitution of non-magnetic La3+ at the Sr2+ 1) could enhance hyperfine fields at 12k site due to strengthening in the Fe3+-O2Fe3+ superexchange interaction giving higher net magnetization. This increase in the hyperfine field of 12k site is evident from the Mossabuer analysis, and 2) may increase the spin canting at the near-neighbor 4f2 site (Sr2+-4f2 ~ 0.366 nm), which results in a positive contribution to the overall magnetic moment. However, the possibility of Co2+ occupying a 4f2 site cannot be ignored, which increases the net positive moment per unit cell of the compound [36] [37]

(a) (b)

Figure 3. Room temperature VSM M vs. H loops of (a) Sr0.7La0.3Fe122xCoxAlxO19 and (b) Sr0.7Pr0.3Fe122xCoxAlxO19.

Table 3. Magnetic parameters of SrFe12O19, Sr0.7La0.3Fe122xCoxAlxO19 and Sr0.7Pr0.3Fe122xCoxAlxO19 measured using VSM. Curie temperature, Tc, of samples was measured using TGA.

Figure 4. Room temperature Ms and Mr of Sr0.7La0.3Fe122xCoxAlxO19 and Sr0.7Pr0.3Fe12-2xCoxAlxO19 measured at 1.2 T field as a function of content, x.

[39]. On the other hand, the substitution of magnetic Pr3+ at the Sr2+ site may not allow spin-canting at the 4f2 site, and with the replacement of Fe3+ at the 2a site, the overall magnetic moment decreases in the Pr3+-SrM compounds.

The variation of coercivity, Hc, of La3+-SrM and Pr3+-SrM compounds as a function of x content is shown in Figure 5. Consistently La3+-SrM samples display higher Hc values than Pr3+-SrM compounds for all x values. The Hc value of La3+-SrM displays a marked ~50% increase at x = 0.2 when compared to that of pure SrFe12O (Hc ~ 3918 Oe). On the other hand, Hc value decreases linearly with increasing the x content for the Pr3+-SrM compound. Being an extrinsic

Figure 5. Coercivity, Hc, of Sr0.7La0.3Fe122xCoxAlxO19 and Sr0.7Pr0.3Fe122xCoxAlxO19 as a function of content, x.

property, the coercivity depends on factors such as crystal structure, morphological features, elemental composition, and defects. In general, Hc can be expressed as,

H c = ( α ( 2 K 1 / ( μ o M s ) ) N ( M r + M s ) / μ o (3)

where α is the microstructure factor which has a reciprocal dependence on the grain size, N is the demagnetization factor, 2K1/(μoMs) is the Ha is the magnetocrystalline anisotropy field [39]. In the M-type hexaferrite, magnetocrystalline anisotropy mainly originates from Fe3+ ions at the 2b and 4f1 site [40] [41] ]. It has been reported that Al3+ ions have a preference for 4f2, 2a, and 12k sites, [42] [43], and Co2+ ions mainly prefer 4f2 (mainly) and 2a sites [37] [38] [44]. In accordance with the Mossbauer analysis, discussed later, it is observed that in both series of samples studied herein, the Co3+-Al3+ ions have a preference for the 2a site. Thus the intrinsic coercivity as per Equation (3), should increase for both series of samples due to the increase in the magnetocrystalline anisotropy, as the site 2b remains unaffected with the said substitution. However, from Figure 5, it is observed that the La3+-SrM sample shows a maximum (5829 Oe) in coercivity value at x = 0.2, followed by a decrease with increasing x content in the sample. On the other hand, the Hc value of the Pr3+-SrM sample shows a monotonous decrease with the x content in the sample. Also, the La3+-SrM sample displays consistently higher coercive field value as compared to Pr3+-SrM. In a comparative study on Sr(La/Pr)Fe12O19 samples it was reported that the Pr3+ doped Sr(Pr)Fe12O19 displayed higher coercivity than La3+ doped Sr(Pr)Fe12O19 sample because of the higher magnetocrystalline anisotropy of the former. However, the contrary results in the present study can be explained by taking microstructure factor, α, in account in Equation (3) above. The La3+-SrM samples have relatively smaller crystallite size as compared to Pr3+-SrM; as a result, the microstructure factor, α, will have a higher value than that of Pr3+-SrM sample. This will make the first term in the Equation (3) above greater than for La3+-SrM as compared to that for Pr3+-SrM. The observed decrease in coercivity with the substitution could be attributed to the fact that the second term in the Equation (3) decreases much rapidly as compared to the αHa term.

The Curie temperature, Tc, of Sr0.7RE0.3Fe12−2xCoxAlxO19 are shown in Figure 6. The maximum Tc value for La3+-SrM and Pr3+-SrM was observed to be 505 and 525 K, respectively at x = 0.40. After attaining maximum value, the Tc decreases gradually for x > 0.40. The overall decrease in Tc after attaining maximum can be explained on the basis of three combined effects [45] [46], 1) substitution of Co2+-Al3+ ions for Fe3+ ions reduce the Fe3+ ions and hence leads to a reduction of Fe3+-O2-Fe3+ number of superexchange interaction and strength and 2) lattice contraction with the substitution alters the bond length and angle of Fe3+-O2-Fe3+ from its optimum interaction strength value. However, in the beginning, Tc drops down for x = 0.20, where presumably Fe3+ might be substituted more with non-magnetic Al3+ and above combined effect become prevalent. However, the increasing magnetic Co2+ ion content helps regain the strength of the superexchange interaction and hence increases the Tc value for x = 0.40 [47].

In order to investigate the site occupancies of transition metal ions and resulting hyperfine parameters in the samples, Mossbauer spectra were collected at room temperature. Figure 7 shows the fitted Mossbauer spectra as a function of content x in the as-synthesized samples. All spectra consist of five Lorentzian sextets originating from 12k, 4f1, 4f2, 2a, and 2b crystallographic sites of Fe3+ ions. The spectra were fitted with the constrain that all the linewidths of the absorption lines were the same. The extracted Mossbauer parameters are listed in Table 4(a) and Table 4(b). The observed sequences of the magnetic hyperfine magnetic fields, HF (4f2 > 2a > 4f1 > 12k > 2b) and isomer shift, δ (the isomer shifts follow the sequence of 4f2 > 12k > 2a > 2b > 4f1) for both samples are in agreement with the reported results [47].

It is well known that the magnitude of the hyperfine magnetic field at the Fe3+ site depends on the distribution of neighboring magnetic cations. The hyperfine field, HF, the parameter of the most intense 12k line effectively shows no variation

Figure 6. Curie temperature, Tc, of Sr0.7La0.3Fe122xCoxAlxO19 and Sr0.7Pr0.3Fe122xCoxAlxO19 as a function of content, x.

(a)(b)

Figure 7. Room temperature fitted Mössbauer spectra of (a) Sr0.7La0.3Fe122xCoxAlxO19 and (b) Sr0.7Pr0.3Fe122xCoxAlxO19 respectively.

(a) (b)

Table 4. (a) Hyperfine parameters of Sr0.7La0.3Fe122xCoxAlxO19 extracted from fitting room temperature Mossbauer spectra; (b) Hyperfine parameters of Sr0.7Pr0.3Fe122xCoxAlxO19 extracted from fitting room temperature Mossbauer spectra.

with x in Pr3+-SrM, while HF increases with x in the La3+-SrM sample. The increase in HF value of La3+-SrM is in agreement with the observed increase in the saturation magnetization, Ms, with content x. Other lines 4f1, 2a, and 4f2 show a monotonous decrease in HF with x.

The isomer shift plot as a function of x is shown in Figure 8. The isomer shift, δ, for both series of samples for all sites visibly remains invariant except for the 2b site. Also, the 2b site’s isomer shift value clearly reflects the influence of the

(a)(b)

Figure 8. Room temperature hyperfine parameters of Sr0.7La0.3Fe122xCoxAlxO19 and Sr0.7Pr0.3Fe122xCoxAlxO19 obtained from Mossbauer spectral analysis.

rare-earth ion in the compound. In the presence of rare-earth ion, the isomer shift value of 2b drops significantly as compared to that of a pure SrFe12O19 compound. Isomer shift is a measure of 3d electron density at the nucleus of the Fe atom. The closest sites of Sr2+ sites are 12k, 4f2, and 2b. The 2b sites are located in the same plane as Sr2+ ions (at a distance of 0.340 nm), while the 4f2 and 12k sites are located in the adjacent planes at nearly the same distance from Sr2+ viz. Sr2+-12k ~ 0.365 nm and Sr2+-4f2 ~ 0.366) nm [48] [49]. Pr [Xe] 4f3 6s2 and La[Xe] 5d1 6s2 with more electrons as compared to Sr (Kr[5s2+]) has a negative effect on the isomer shift, that is increased substitution of Pr3+ or La3+ increases the net s-electron charge density at the 2b site, which in turn decreases the isomer shift value at the 2b site.

The quadrupole shift, QS, values of 2b site show a similar ~10% drop in both Pr3+ and La3+ substituted samples. The quadrupole values for all sites remain invariant with x for both series of samples. It is known that QS is a measure of the asymmetry of a crystallographic site. Since the 2b sites are located in the same plane as Sr2+, the substitution of the smaller Pr3+ and La3+ for Sr2+ ions induces a deformation, which affects the electronic site symmetry of the nearby 2b site. Due to the substitution of smaller La3+ and Pr3+ ion for Sr2+ ion, the lattice constant c contracts. Consequently, large Fe-O distances parallel to the c-axis decreases. As a result, the oxygen bipyramid of the 2b-site becomes more symmetric, which is in agreement with the observed reduction in the quadrupole shift value with content x. A similar decrease in 2b quadrupole splitting originating from the presence of La3+ substitution has been reported in Sr1xLaxFe12yCoyO19 [39].

A close examination of the fitted spectra shows a decreasing trend of 2a absorption intensities with increasing x. Without Co2+-Al3+ co-doping, the intensity of 2a site lines in both Sr0.7La0.3Fe12O19 and Sr0.7Pr0.3Fe12O19 are nearly the same as that of undoped SrFe12O19. However, with the substitution of Co2+-Al3, the 2a absorption area decreased linearly and finally disappeared at x = 0.6 and 1.0 for La3+-SrM and Pr3+-SrM sample, respectively. This suggests preferential occupation of Co2+-Al3 for the 2a site. It has already been reported that Al3+ ion prefers 12k and 2a site [50] while Co2+ was reported to enter 4f2 site in Sr1xLaxFe12xCoxO19 [51]. There is only one 2a site in every one formula unit of M-type hexaferrite. Therefore, the fact that the 2a absorption line of Sr0.7La0.3Fe122xCoxAlxO19 disappeared completely at x = 0.6 means that all the Co2+-Al3 dopants almost exclusively go into 2a site. In the case when x > 0.5, the excessive Co2+-Al3 dopant should occupy the crystallographic sites other than the 2a site. On the other hand, in the case of Sr0.7Pr0.3Fe122xCoxAlxO19, 2a absorption intensity has decreased linearly and has completely diminished at x = 1.0, which indicates that all the 2a sites become full when x = 1.0. This suggests that only half of the Co2+-Al3+ dopant occupies the 2a site compared to the case of the La3+-SrM sample. Then, the other half of the 2a site should be occupied by Fe3+ ions through the range of x examined. In the case of Pr3+ series samples, a slight decline in the absorption area for the most intense 12k site is observed, while in the La3+ containing samples, 12k areal almost remains constant. Thus it seems that the excessive Co2+-Al3+ dopant occupies the 2a site and 12k site in Sr0.7Pr0.3Fe122xCoxAlxO19. The above observation clearly indicates that the tendency of 12k site occupation by the Co2+-Al3+ dopant is affected by the species of rare-earth ions at the Sr2+ site. Where Co2+-Al3+ impurity mostly prefers 2a site and its preference for 12k site is enhanced in the presence of Pr3+ rather than La3+, once the 2a site is completely filled. This observation is interesting considering the fact that there are no 2a sites in the vicinity of Sr2+, and hence the substitution at Sr2+ with rare-earth ions is not expected to influence the site occupancy at the 2a site. Furthermore, the physical reason for the preferential site occupancy of Co2+-Al3+ in the presence of La3+ or Pr3+ is not clear at this stage. Mossbauer spectroscopy cannot discriminate against the site occupancies of both Co2+ and Al3+ ions separately. But one scenario is that Co2+ ions might not prefer the 2a site in the presence of Pr3+ ion, which needs further investigation.

4. Conclusion

The structural, magnetic, and Mossbauer spectrum study of the influence of rare-earth La3+ and Pr3+ on Sr0.7RE0.3Fe122xCoxAlxO19, (RE: La and Pr) hexaferrite compounds were investigated. Overall, dopant ion brings a unit cell contraction, which affects the Curie temperature and saturation magnetization of both series of samples. La3+-SrM samples show marked improvement in the magnetization value, while Pr3+-SrM samples show deterioration in Ms value with the Co2+-Al3+ substitution. This is suggestive of the fact that magnetic Pr3+, unlike non-magnetic La3+ (no 4f electrons), substitution for Sr2+ could be detrimental to the overall magnetization behavior of the SrFe12O19 compound. The Co2+-Al3+ substitution brings in grain refinement, more for the La3+-SrM than Pr3+-SrM, which has a considerable impact on the coercivity values of the compound. The Mossbauer spectral analysis revealed 2a as a preferred substitution for Co2+-Al3+ ions for both series of samples. Due to lattice contraction, the 2b site displayed a considerable reduction in its quadrupole shift value. The improved hyperfine field for the 12k site in the La3+-SrM series is attributed to the increase in its magnetization value with the substitution of Co2+-Al3+ ions for Fe3+. For the study, it is concluded that the substitution of rare-earth ions with 4f charge distribution for Sr2+ could be detrimental to the overall magnetic performance of the M-type hexaferrite materials.

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

References

[1] Chen, N., Yang, K. and Mingyuan, G. (2010) Microwave Absorption Properties of La-Substituted M-Type Strontium Ferrites. Journal of Alloys and Compounds, 490, 609-612.
https://doi.org/10.1016/j.jallcom.2009.10.116
[2] Riches, E.E. (1972) Ferrites: A Review of Materials and Applications. 3rd Edition, Mills & Boon, Richmond.
[3] Chin, T.-S. (2000) Permanent Magnet Films for Applications in Microelectromechanical Systems. Journal of Magnetism and Magnetic Materials, 209, 75-79.
https://doi.org/10.1016/S0304-8853(99)00649-6
[4] Hernandez, P., De Francisco, C., Munoz, J.M., Iniguez, J., Torres, L. and Zazo, M. (1996) Influence of Sintering Atmosphere on the Magnetic After-Effect in Strontium Ferrites. Journal of Magnetism and Magnetic Materials, 157-158, 123-124.
https://doi.org/10.1016/0304-8853(95)01070-X
[5] Jin, Z.Q., Tang, W., Zhang, J.R., Lin, H. and Du, Y.W. (1998) Magnetic Properties of Isotropic SrFe12O19 Fine Particles Prepared by Mechanical Alloying. Journal of Magnetism and Magnetic Materials, 182, 231-237.
https://doi.org/10.1016/S0304-8853(97)00679-3
[6] Qiao, L., You, L., Zheng, J., Jiang, L. and Sheng, J. (2007) The Magnetic Properties of Strontium Hexaferrites with La-Cu Substitution Prepared by SHS Method. Journal of Magnetism and Magnetic Materials, 318, 74-78.
https://doi.org/10.1016/j.jmmm.2007.04.028
[7] Besenicar, S. and Miha, D. (1991) High Coercivity Sr Hexaferrites. Journal of Magnetism and Magnetic Material, 101, 307-309.
https://doi.org/10.1016/0304-8853(91)90763-Z
[8] Rezlescu, L., Rezlescu, E., Popa, P.D. and Rezlescu, N. (1999) Fine Barium Hexaferrite Powder Prepared by the Crystallization of Glass. Journal of Magnetism and Magnetic Materials, 193, 288-290.
https://doi.org/10.1016/S0304-8853(98)00442-9
[9] Zhong, W., Ding, W.P., Zhang, N., Hong, J.M., Yan, Q.J. and Du, Y.W. (1997) Key Step in Synthesis of Ultrafine BaFe12O19 by Sol-Gel Technique. Journal of Magnetism and Magnetic Materials, 168, 196-202.
https://doi.org/10.1016/S0304-8853(96)00664-6
[10] Kojima, H. (1982) Chapter 5. Fundamental Properties of Hexagonal Ferrites with Magnetoplumbite Structure. In: Handbook of Ferromagnetic Materials, Elsevier, Amsterdam, Vol. 3, 305-391.
https://doi.org/10.1016/S1574-9304(05)80091-4
[11] Luo, H., Rai, B.K., Mishra, S.R., Nguyen, N.N. and Liu, J.P. (2012) Physical and Magnetic Properties of Highly Aluminum Doped Strontium Ferrite Nanoparticles Prepared by Auto-Combustion Route. Journal of Magnetism and Magnetic Materials, 324, 2602-2608.
https://doi.org/10.1016/j.jmmm.2012.02.106
[12] Shirtcliffe, N.J., Thompson, S., O’keefe, E.S. Appleton, S. and Perry, C.C. (2007) Highly Aluminium Doped Barium and Strontium Ferrite Nanoparticles Prepared by Citrate Auto-Combustion Synthesis. Materials Research Bulletin, 42, 281-287.
https://doi.org/10.1016/j.materresbull.2006.06.001
[13] Fang, Q.Q., Liu, Y.M., Yin, P. and Li, X.G. (2001) Magnetic Properties and Formation of Sr Ferrite Nanoparticle and Zn, Ti/Ir Substituted Phases. Journal of Magnetism and Magnetic Materials, 234, 366-370.
https://doi.org/10.1016/S0304-8853(01)00428-0
[14] Iqbal, M.J., Ashiq, M.N., Gomes, P.H. and Munoz, J.M. (2008) Synthesis, Physical, Magnetic and Electrical Properties of Al-Ga Substituted Co-Precipitated Nanocrystalline Strontium Hexaferrite. Journal of Magnetism and Magnetic Materials, 320, 881-886.
https://doi.org/10.1016/j.jmmm.2007.09.005
[15] Albanese, G., Carbucicchio, M. and Deriu, A. (1974) Temperature Dependence of the Sublattice Magnetizations in Al- and Ga-Substituted M-Type Hexagonal Ferrites. Physica Status Solidi, 23, 351-358.
https://doi.org/10.1002/pssa.2210230202
[16] Ashiqa, M.N., Iqbalb, M.J. and Gulc, I.H. (2009) Structural, Magnetic and Dielectric Properties of Zr-Cd Substituted Strontium Hexaferrite (SrFe12O19) Nanoparticles. Journal of Alloys and Compounds, 487, 341-345.
https://doi.org/10.1016/j.jallcom.2009.07.140
[17] Le Breton, J.M., Teillet, J., Wiesinger, G., Morel, A., Kools, F. and Tenaud, P. (2002) Mossbauer Investigation of Sr-Fe-O Hexaferrites with La-Co Addition. IEEE Transactions on Magnetics, 38, 2952-2954.
https://doi.org/10.1109/TMAG.2002.803177
[18] Nourbakhsh, A.A., Noorbakhsh, M., Nourbakhsh, M., Shaygan, M. and Mackenzie, K.J.D. (2011) The Effect of Nano Sized SrFe12O19 Additions on the Magnetic Properties of Chromium-Doped Strontium-Hexaferrite Ceramics. Journal of Materials Science: Materials in Electronics, 22, 1297-1302.
https://doi.org/10.1007/s10854-011-0303-3
[19] Liu, X.S., Zhong, W., Yang, S., Yu, Z., Gu, B.X. and Du, Y. (2002) Structure and Magnetic Properties of La3+-Substituted Strontium Hexaferrite Particles Prepared by Sol-Gel Method. Physica Status Solidi, 193, 314-319.
https://doi.org/10.1002/1521-396X(200209)193:2<314::AID-PSSA314>3.0.CO;2-W
[20] Dahal, J.N., Neupane, D. and Mishra, S.R. (2019) Exchange-Coupling Behavior in SrFe12O19/La0.7Sr0.3Mno0.3 Nanocomposites. Ceramics, 2, 100-111.
https://doi.org/10.3390/ceramics2010010
[21] Mocuta, H., Lechevallier, L., Le Breton, J.M., Wang, J.F. and Harris, I.R. (2004) Structural and Magnetic Properties of Hydrothermally Synthesised Sr1-xNdxFe12O19 Hexagonal Ferrites. Journal of Alloys and Compounds, 364, 48-52.
https://doi.org/10.1016/S0925-8388(03)00545-0
[22] Lechevallier, L., Le Breton, J.M., Wang, J.F. and Harris, I.R. (2004) Structural and Magnetic Properties of Hydrothermally Synthesised Sr1-xNdxFe12O19 Hexagonal Ferrites. Journal of Magnetism and Magnetic Materials, 269, 192-196.
https://doi.org/10.1016/S0304-8853(03)00591-2
[23] Wang, J.F., Ponton, C.B. and Harris, I.R. (2001) A Study of the Magnetic Properties of Hydrothermally Synthesised Sr Hexaferrite with Sm Substitution. Journal of Magnetism and Magnetic Materials, 234, 233-240.
https://doi.org/10.1016/S0304-8853(01)00366-3
[24] Lechevallier, L., Le Breton, J.M., Morel, A. and Tenaud, P. (2008) On the Solubility of Rare Earths in M-Type SrFe12O19 Hexaferrite Compounds. Journal of Physics: Condensed Matter, 20, Article ID: 175203.
https://doi.org/10.1088/0953-8984/20/17/175203
[25] Polyko, D.D., Bashkirov, L.A., Trukhanov, S.V., Lobanovskii, L.S. and Sirota, I.M. (2011) Crystal Structure and Magnetic Properties of High-Coercivity Sr1-xPrx Fe12-xZnxO19 Solid Solutions. Inorganic Materials, 47, 77-79.
https://doi.org/10.1134/S0020168511010110
[26] Yang, Y.J. and Liu, X.S. (2014) Microstructure and Magnetic Properties of La-Cu Doped M-Type Strontium Ferrites Prepared by Ceramic Process. Materials Technology, 29, 232-236.
[27] Corral Huacuz, J.C. and Mendoza-Suárez, G. (2002) Preparation and Magnetic Properties of Ir-Co and La-Zn Substituted Barium Ferrite Powders Obtained by Sol-Gel. Journal of Magnetism and Magnetic Materials, 430, 430-433.
https://doi.org/10.1016/S0304-8853(01)01141-6
[28] Torkian, S., Ghasemi, A., Shojarazavi, R. and Tavoosi, M. (2016) Structural and Magnetic Properties of High Coercive Al-Substituted Strontium Hexaferrite Nanoparticles. Journal of Superconducativity and Novel Magnetism, 29, 1627-1640.
https://doi.org/10.1007/s10948-016-3450-1
[29] Trusov, L.A., Gorbachev, E.A., Lebedev, V.A., Sleptsova, A.E., Roslyakov, I.V., Kozlyakova, E.S., Vasiliev, A.V., Dinnebier, R.E. Jansen, M. and Kazin, P.E. (2018) Ca-Al Double-Substituted Strontium Hexaferrites with Giant Coercivity, Chemical Communications, 54, 479-482.
https://doi.org/10.1039/C7CC08675J
[30] Narang, S.B., Singh, A. and Singh, K. (2014) High Frequency Dielectric Behavior of Rare Earth Substituted Sr-M Hexaferrite. Journal of Ceramic Processing Research, 8, 347.
[31] Shannon, R.D. (1976) Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallographica, A32, 751-767.
https://doi.org/10.1107/S0567739476001551
[32] Wagner, T.R. (1998) Preparation and Crystal Structure Analysis of Magnetoplumbite-Type Baga12O19. Journal of Solid State Chemistry, 136, 120-124.
https://doi.org/10.1006/jssc.1997.7681
[33] Cullity, B.D. (1956) Elements of X-Ray Diffraction. 2nd Edition, Addison-Wesley Publishing, Boston.
[34] Dahal, J.N., Wang, L., Mishra, S.R., Nguyen, V.V., and Liu, J.P. (2014) Synthesis and Magnetic Properties of SrFe12-x-yAlxCoyO19 Nanocomposites Prepared via Autocombustion Technique. Journal of Alloys and Compounds, 595, 213-220.
https://doi.org/10.1016/j.jallcom.2013.12.186
[35] Zi, Z.F., Sun, Y.P., Zhu, X.B., Yang, Z.R. and Song, W.H. (2008) Structural and Magnetic Properties of SrFe12O19 Hexaferrite Synthesized by a Modified Chemical Co-Precipitation Method. Journal of Magnetism and Magnetic Materials, 320, 2746-2751.
https://doi.org/10.1016/j.jmmm.2008.06.009
[36] Wiesinger, G., Muller, M., Grossinger, R., Pieper, M., Morel, A., Kools, F., Tenad, P., Le Breton, J.M. and Kreisel, J. (2002) Substituted Ferrites Studied by Nuclear Methods. Physica Status Solidi, 189, 499-508.
https://doi.org/10.1002/1521-396X(200202)189:2<499::AID-PSSA499>3.0.CO;2-H
[37] Lechevallier, L., Le Breton, J.M., Teillet, J. Morel, A., Kools, F. and Tenaud, P. (2003) Mossbauer Investigation of Sr1-xLaxFe12-yCoyO19 Ferrites. Physica B: Condensed Matter, 327, 135-139.
https://doi.org/10.1016/S0921-4526(02)01712-X
[38] Le Breton, J.-M., Teillet, J., Wiesinger, G., Morel, A., Kools, F. and Tenaud, P. (2002) Mossbauer Investigation of Sr-Fe-O Hexaferrites La-Co Addition. IEEE Transactions on Magnetics, 38, 2952-2954.
https://doi.org/10.1109/TMAG.2002.803177
[39] Kools, F. ((1985) Factors Governing the Coercivity of Sintered Anisotropic M Type Ferrite. Le Journal de Physique Colloques, 46, 346-349.
[40] Zhou, X.Z., Morrish, A.H., Li, Z.W. and Hong, Y.K. (1991) Site Preference for Co2+ and Ti4+ in Co-Ti Substituted Barium Ferrites. IEEE Transactions on Magnetics, 27, 4654-4656.
https://doi.org/10.1109/20.278906
[41] Han, D.H., Yang, Z., Zeng, H.X., Zhou, X.Z. and Morrish, A.H. (1994) Cation Site Preference and Magnetic Properties of Co-Sn-Substituted Ba Ferrites Particles. Journal of Magnetism and Magnetic Materials, 137, 191-196.
https://doi.org/10.1016/0304-8853(94)90205-4
[42] Choi, D.H., An, S.Y., Lee, S.W., Shim, I.-B. and Kim, C.S. (2004) Site Occupancy and Anisotropy Distribution of Al Substituted Ba-Ferrite with High Coercivity. Physica State Solid, 241, 1736-1739.
https://doi.org/10.1002/pssb.200304633
[43] Albanese, G. (1995) Mossbauer Investigation of Aluminium Substituted Barium Hexaferrites in the Paramagnetic State. Journal of Magnetism and Magnetic Materials, 147, 421-426.
https://doi.org/10.1016/0304-8853(95)00063-1
[44] Lechevallier, L., Le Breton, J.M., Morel, A. and Tenaud, P. (2007) Influence of the Presence of Co on the Rare Earth Solubility in M-Type Hexaferrite Powders. Journal of Magnetism and Magnetic Materials, 316, e109-e111.
https://doi.org/10.1016/j.jmmm.2007.02.042
[45] Ghimire, M., Yoon, S., Wang, S.L., Neupane, D., Alam, J. and Mishra, S.R. (2018) Influence of La Content on Magnetic Properties of Cu Doped M-Type Strontium Hexaferrite: Structural, Magnetic, and Mossbauer Spectroscopy Study. Journal of Magnetism and Magnetic Materials, 454, 110-120.
https://doi.org/10.1016/j.jmmm.2018.01.062
[46] Kunwar, D.L., Neupane, D., Dahal, J.N. and Mishra, S.R. (2019) Structural, Magnetic, and Electrical Properties of Re Doped Sr0.82Re0.18Fe12-xAlxO19 (Re = Gd, Pr, Sm) Compound. Advances in Materials Physics and Chemistry, 9, 175.
https://doi.org/10.4236/ampc.2019.99014
[47] Evans, B.J., Grandjean, F., Lilot, A.P., Vogel, R.H. and Gerard, A. (1987) 57Fe Hyperfine Interaction Parameters and Selected Magnetic Properties of High Purity MFe12O19 (M = Sr, Ba). Journal of Magnetism and Magnetic Materials, 67, 123-129.
https://doi.org/10.1016/0304-8853(87)90728-1
[48] Wartewig, P., Krause, M.K., Esquinazi, P., Rosler, S. and Sonntag, R. (1999) Magnetic Properties of Zn- and Ti-Substituted Barium Hexaferrite. Journal of Magnetism and Magnetic Materials, 192, 83-99.
https://doi.org/10.1016/S0304-8853(98)00382-5
[49] Obradors, X., Solans, X.A., Collomb, D., Samaras, J., Rodriguez, M., Pernet, M. and Font-Altaba, M. (1988) Crystal Structure of Strontium Hexaferrite SrFe12O19. Journal of Solid State Chemistry, 72, 218-224.
https://doi.org/10.1016/0022-4596(88)90025-4
[50] Dixit, V., Nandadasa, C.N., Kim, S.G., Kim, S., Park, J., Hong, Y.K. and Moitra, A. (2015) Site Occupancy and Magnetic Properties of Al-Substituted M-Type Strontium Hexaferrite. Journal of Applied Physics, 117, Article ID: 243904.
https://doi.org/10.1063/1.4922867
[51] Lechevallier, L. and Le Breton, J.M. (2005) Substitution Effects in M-Type Hexaferrite Powders Investigated by Mossbauer Spectrometry. Journal of Magnetism and Magnetic Materials, 290, 1237-1239.
https://doi.org/10.1016/j.jmmm.2004.11.411

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.