Tuning of magnetic properties of aluminium-doped strontium hexaferrite powders
Ma Xiao-Mei, Liu Jie, Zhu Sheng-Zhi, Shi Hui-Gang†,
Key Laboratory for Magnetism and Magnetic Materials of the Ministry of Education, Lanzhou University, Lanzhou 730000, China

 

† Corresponding author. E-mail: shihuig@lzu.edu.cn

Abstract
Abstract

M-type Al-doped strontium ferrite powders (SrAlxFe2nxO19, n = 5.9) with nominal Al content of x = 0–2.0 are prepared by traditional ceramic technology. The phase identification of the powders, performed using x-ray diffraction, shows the presence of purity hexaferrite structure and absence of any secondary phase. The lattice parameters decrease with increasing x. The average grain size of the powders is about 300 nm–400 nm at Al3+ ion content x = 0–2.0. The room-temperature hysteresis loops of the powders, measured by using vibrating sample magnetometer, show that the specific saturation magnetization (σs) value continuously decreases while the coercivity (Hc) value increases with increasing x, and Hc reaches to 9759 Oe (1 Oe = 79.5775 A/m) at x = 2.0. According to the law of approach saturation, Hc value increases with increasing Al3+ ion content, which is attributed to the saturation magnetization (Ms) decreasing more rapidly than the magnetic anisotropy constant (K1) obtained by numerical fitting of the hysteresis loops. The distribution of Al3+ ions in the hexaferrite structure of SrAlxFe2nxO19 is investigated by using 57Co Mössbauer spectroscopy. The effect of Al3+ doping on static magnetic properties contributes to the improvement of magnetic anisotropy field.

1. Introduction

M-type hexagonal ferrite (MFe12O19, M is divalent cation, such as Ba2+, Sr2+, and so on), is one of the most used hard magnetic materials with applications such as permanent magnets, magnetic recording media due to their low cost, good chemical stabilities, high coercivities, high saturation magnetizations, and large magnetocrystalline anisotropies.[15] Recently, the barium ferrite was widely used as magnetic recording media, microwave devices, and electromagnetic wave absorbers.[3,6,7] Furthermore, various methods of fabricating the M-type hexaferrite have been developed including sol-gel method,[8,9] co-precipitation,[10,11] ceramic method,[12,13] and hydrothermal method,[14] citrate auto-combustion,[15] and glycin–nitrate method.[16]

In order to change physical properties, substituting some cations into SrM structure is an effective method to cope with the different applications. A lot of work has been done to modify the magnetic parameters of SrM by substituting Fe3+ with other cations, such as Al3+, Cr3+, Ga3+, and so on.[1719] The magnetic properties of substituted hexaferrites directly depend on the electronic configuration of the dopant cations as well as their preference to occupy the different Fe sub-lattices of the magnetoplumbite structure.[20] In particular, as is well known, Al-substituted SrM has very large coercivity.[17,21,22] Furthermore, it is important to understand the mechanism for increasing the coercivity by Al3+ doping.

SrM with hexagonal structure and molecular formula SrFe12O19 has a magnetoplumbite (M) structure which was determined by Adelsklod.[23,24] The crystal structure can be separated into two blocks, one block is R Block and another one is S block , R*S* block can be obtained by rotating the corresponding block 180° around the hexagonal c axis, so the chemical composition of the unit cell is Sr2Fe24O38. The 24 Fe3+ ions of the unit cell are among five different crystal sites, i.e., three octahedral sites (12k, 4f2, and 2a), one tetrahedron site (4f1) and one trigonal bipyramidal site (2b).[25] The 16 Fe3+ ions with upward spins are located at 2a, 2b, and 12k, whereas the remaining 8 Fe3+ ions are located at 4f1 and 4f2 with downward spins, which result in a net magnetic moment of 40 μB per unit cell.[26] It has large magnetocrystalline anisotropy by the strong exchange interaction among the Fe3+ ions. Therefore the substitution for Fe ions could be an effective way to modify the magnetic properties.

In this study, we prepare Al3+-doped SrM by traditional ceramic technology, and the coercivity reaches to 9759 Oe at x = 2.0. We attempt to understand in detail the reasons and the mechanism of enhancing the coercivity and magnetic structure formation of SrAlxFe2nxO19 as a function of the substituted amounts of dopants in a range (0 ≤ x ≤ 2.0).

2. Experiment

SrAlxFe2nxO19 (n = 5.9 x = 0–2.0) samples were prepared via ball milling appropriate mixtures of SrCO3, Fe2O3 and Al2O3 powders. The amount of each compound was determined for each particular stoichiometry of SrAlxFe2nxO19 and then mixed in a ball milling cylinder. The mechanical alloying was carried out in a traditional ball mill for 4 h and the volume ratio of alcohol to powders was about 1.5:1. After milling, the powders were placed into cylindrical pellets. Calcination was then carried out for 2 h in air at 1200 °C.

The phase composition and crystal structure were analyzed by using Bruker diffractometer with Cu Kα x-ray diffraction measurement (λ = 1.5406 Å); scanning electron microscopy was employed to analyze the morphologies and microstructures of the samples. Magnetic properties could be obtained by analyzing the data of the vibrating sample magnetometer and the Mössbauer spectroscopy. The Mössbauer spectrum at room temperature was recorded using a Mössbauer spectrometer operating in constant acceleration mode (triangular wave) in transmission geometry. The source employed was 57Co in Rh matrix of strength 50 mCi. The velocity was scaled by using an enriched a-Fe metal foil under an effective nuclear hyperfine field (Heff) of 331 kOe at room temperature.

3. Result and discussion

Figure 1 shows the typical x-ray diffraction (XRD) patterns of the precursor powders without and with various Al concentrations (x = 0–2.0) annealed for 2 h at 1200 °C. All the XRD patterns are similar, and consist of a set of strong diffraction profiles, which are similar to those of the SrFe12O19 (ICCD card No. 79-1411). There is no impurity phase (no additional minor peaks can be observed), thus confirming the formation of pure and well-crystallized phase. It is found that the peak positions systematically shift towards the high values with increasing doping concentration, which shows that there appears to be some change in crystal structure as x increases.

Fig. 1. XRD patterns of SrAlxFe2nxO19 at T = 1200 °C.

The lattice parameters a and c of SrAlxFe2nxO19 (n = 5.9, x = 0–2.0) are calculated according to the following formula:

where h, k, and l are Miller indices and d is the interplanar spacing as determined by the Bragg formula, 2d sin θ = . The unit cell volume is calculated by using the lattice parameters a and c from the following formula:[27,28]

The x-ray density, ρx, of the material is calculated according to the following relation:

where M represents the molar mass of the sample and Na is the Avogadro’s number.

The changes of structural parameters, density and volume with composition are given in Table 1. It can be observed that the values of lattice parameters a and c decrease with x increasing from 5.89 Å to 5.84 Å and from 23.11 Å to 22.89 Å, respectively. This decrease becomes slow with composition rising up to x = 1.4 and then decreases dramatically for x = 2.0. As a result, the volume also decreases, which is attributed to the radius of Al3+ (0.535 Å) being smaller than that of Fe3+ (0.645 Å).[29] Furthermore, the density of samples decreases with increasing x.

Table 1.

Values of crystal lattice a and c, the unit-cell volume and the density for SrAlxFe2nxO19.

.

Particle size and morphology are studied by SEM as shown in Figs. 2(a)2(d). It is noteworthy that there is nearly no change in grain size (300 nm to 400 nm) nor morphology in the range of Al3+ substitution from x = 0 to x = 2.0 at the same temperature. It could be considered that Al3+ substitution does not affect the particle size and morphology, but it makes great contributions to the lattice parameters and the magnetic properties. The influence on the magnetic properties will be discussed in detail later.

Fig. 2. Scanning electron micrographs of SrAlxFe2nxO19 magnetic powder particles with (a) x = 0, (b) x = 0.8, (c) x = 1.4, (d) x = 2.0.

In the hexagonal structure of strontium ferrite, Fe3+ ions occupy five sites: 4f1, 2a, 4f2, 12k, and 2b. The contributions of Fe3+ ions located on those sites to magnetic properties are different from each other. After Al3+ ions enter into the lattice and substitute for some Fe3+ ions, the magnetocrystalline anisotropy field is changed. From the Mössbauer spectra as shown in Fig. 3, the substitution sites can be investigated, and then the effect of substitution on anisotropy field can be revealed. The hyperfine parameters including hyperfine field, isomer shift, quadrupole splitting and Fe3+ occupation area ratio are shown in Figs. 47. It can be observed in Fig. 4 that the isomer shifts of octahedral sites (12k, 4f2, and 2a) are greater than those of tetrahedron sites, because the electron density of Fe3+ ions at octahedral sites is smaller than that at the octahedral sites. The isomer shift value in the range 0.24 mm/s–0.40 mm/s for five sites at x = 0–2.0 implies only the existence of high-spin Fe3+ cations.[30,31] In Fig. 5, the hyperfine fields of five sites are in the following sequence: Hhf (2b) < Hhf (12k) < Hhf (4f1) < Hhf (2a) < Hhf (4f2), which is in agreement with the results reported by Evan et al.[32] The hyperfine field of each site decreases only slightly with increasing x, because the magnetic Fe3+ ions are replaced by non-magnetic Al3+ ions. The numbers of cations located at sites 12k, 4f1, 4f2, 2a, and 2b are 6, 2, 2, 1, and 1 (per formula unit), and the theoretical areas of the sextets are 50, 16.7, 16.7,8.3, and 8.3, respectively. The sextet area is directly proportional to the number of Fe3+ ions in the corresponding sites.[30] In the present study, the corresponding values of relative areas for x = 0 are found to be 51, 11.2, 8.2, 22.4, and 7.2, which is caused by the overlapping of the spectral peaks. From Fig. 6, Al3+ mainly replaces Fe3+ on 2a and 4f1 sites for x ≤ 1.4 as reported in Ref. [6]. For x = 2.0, the area of 12k sites has a sharp decline in the relative strength,which implies that Al3 + ions are more inclined to replace Fe3+ of 12k site than Fe3+ of 4f1 site. These results indicate that the substitution Al3+ ions occur mainly at 2a, 4f1, and 12k sites, and Fe3 + of 2b site can hardly be substituted for x ≤ 2.0. In Fig. 7, the quadrupole splitting value of 2b site is very large as compared with those of other sites, which is due to the strongly distorted environment around 2b site.[25,33] After Al3+ ions enter into the lattice and substitute for some Fe3+ ions, Mössbauer parameters of SrAlxFe2−xO19 at room temperature are shown in Table 2.

Fig. 3. The room temperature Mössbauer spectra of SrAlxFe2nxO19 with (a) x = 0, (b) x = 0.8, (c) x = 1.4, and (d) x = 2.0.
Fig. 4. Isomer shifts on five sites of SrAlxFe2nxO19.
Fig. 5. Hyperfine fields on five sites of SrAlxFe2nxO19.
Fig. 6. Fe3+ ion occupation area ratio on five sites of SrAlxFe2nxO19.
Fig. 7. Plots of quadrupole splitting versus Al doping amount on five sites of SrAlxFe2nxO19.
Table 2.

Mössbauer parameters of SrAlxFe2nxO19 at room temperature.

.

The hysteresis loops of powder samples are measured at room temperature by using a vibrating sample magnetometer in Fig. 8. The Hc can be directly obtained from the loop. The magnetocrystalline anisotropy constant (K1) and σs can be obtained by using the law of approach to saturation which is the sole static method to determine them for non-oriented polycrystalline materials. For the stage of approach to saturation in the hysteresis loop, the magnetization behavior can be described as[3436]

The A/H term is related to the existence of homogeneity in a microcrystal and is often ignored in the higher field case. The B/H2 is related to the magnetic anisotropy constant.[33] The χ is the high field susceptibility. For hexagonal ferrite, B can be defined as[36]

Fig. 8. MH loops for SrAlxFe2nxO19 at 1200 °C.

The saturation magnetization Ms and magnetic anisotropy constant (K1) can be obtained by fitting the hysteresis loop using Eqs. (4) and (5) as shown in Table 3. It is noticeable that the value of Ms decreases with increasing Al3+ doping, while the coercivity increases.

Table 3.

The room temperature magnetic parameters of Al3+ substituted SrAlxFe2nxO19.

.

The coercivity of M-type ferrite can be expressed as[37]

Here, there are two kinds of anisotropies HA and HD: HA is magnetic anisotropy field and HD is a shape-demagnetizing anisotropy field. Furthermore HD, which varies only with the shape of the particles, lies within a very small range (as the results show in Fig. 2). Therefore, the origin of the coercivity is derived from the magnetic anisotropy field HA, while HD can be ignored. HA can be given by[38]

In Fig. 9, it is obvious that the saturation magnetization Ms decreases more rapidly than the magnetic anisotropy constant (K1) with the increase of Al3+ ion content, which can cause the magnetic anisotropy field to increase. As for the results of Mössbauer spectra, the reduction of saturation magnetization is attributed to the fact that the Al3+ ions preferentially occupy 2a, 4f1, and 12k sites for x ≤ 2.0, and a large number of Fe3+ on 12k sites, which are replaced by Al3+ contributed to the saturation magnetization, drastically decrease for x = 2.0. In addition, super-exchange interaction of Fe3+–O2−–Fe3+ is also among the factors that can influence the magnetic properties. Super-exchange interaction decreases with non-magnetic Al3+ replacing magnetic Fe3+ sites that can lead to the decrease of saturation magnetization. Furthermore, Xu et al.[39] found that the Fe3+ ions on 12k sites make a negative contribution to magnetocrystalline anisotropy constant, whereas Fe3+ ions on 2b sites result in a large positive contribution.[40] In this work, Al3+ ions mainly replace Fe3+ ions of 2a, 4f1, and 12k sites and Fe3+ ions on 2b sites can hardly be substituted by Al3+ ions, as evidenced from the Mössbauer results for x ≤ 2.0, which also causes the enhancement of magnetic anisotropy field.

Fig. 9. Dependencies of saturation magnetization, Ms, and magnetic anisotropy constant, K1, on Al3+ ion substitution in SrAlxFe2nxO19.
4. Conclusions

A series of SrAlxFe2nxO19 powders (n = 5.9, 0 ≤ x ≤ 2.0) are synthesized by conventional ceramic method. At low substitutions, Al3+ ions mainly occupy the 2a, 4f1, and 12k sites of Fe3+. At x = 2.0, the decrease of magnetizations is attributed to the fact that Fe3+ on 2b sites can hardly be substituted. The magnetic moment per unit cell, saturation magnetization and magnetic anisotropy constant decrease with increasing Al3+ substitution, while the coercivity increases. The enhancement of coercivity is caused by increasing magnetic anisotropy field, which is attributed to the fact that the saturation magnetization (Ms) decreases more rapidly than the magnetic anisotropy constant (K1) with the increase of Al3+ substitution.

Reference
1Zou HLi S HZhang L QYan S HWu H GZhang STian M 2011 J. Magn. Magn. Mater. 323 1643
2Nie YHarward IBalin KBeaubien ACelinski Z 2010 J. Appl. Phys. 107 073903
3Dehlinger A SLeBerre MCanut BChatelon J PAlbertini DPerrot SGivord DRousseau J J 2010 J. Magn. Magn. Mater. 322 3293
4Luo J HXu YMao H K 2015 J. Magn. Magn. Mater. 381 365
5Li T TLi YWu R NZhou HFang X CSu S BXia A LJin C GLiu X G 2015 J. Magn. Magn. Mater. 393 325
6Qiu J XZhang Q GGu M Y 2005 J. Appl. Phys. 98 103905
7Meng P YXiong KJu KLi S NXu G L 2015 J. Magn. Magn. Mater. 385 407
8Sapoletova N AKushnir S ELi Y HAn S YSeo J WHur K H 2015 J. Magn. Magn. Mater. 389 101
9Pawar R ADesai S STamboli Q YShirsath S EPaatange S M 2015 J. Magn. Magn. Mater. 378 59
10Alam R SMoradi MRostami MNikmannesh HMoayedi RBai Y 2015 J. Magn. Magn. Mater. 381 1
11Chin T SHsu S LDeng M C 1993 J. Magn. Magn. Mater. 120 64
12Lotgering F K1974J. Phys & Chem. Solid351633
13Van Uiteit L G 1957 J. Appl. Phys. 28 317
14Pankov V VPernet MGermi PMollard P 1993 J. Magn. Magn. Mater. 120 69
15Shirtcliffea N JThompson SO’Keefe E SAppleton SPerry C C 2007 Mater. Res. Bull. 42 281
16Wang H ZYao BXu YHe QWen G HLong S WFan JLi G DShan LLiu BJiang L NGao L L 2012 J. Alloys Compd. 537 43
17Luo HRai B KMishra S HNguyen V VLiu J P 2012 J. Magn. Magn. Mater. 324 2602
18Kazina P ETrusovb L AZaitsevb D DTretyakova Y DJansenc M 2008 J. Magn. Magn. Mater. 320 1068
19Awawdeh MBsoul IMahmood S H 2014 J. Alloys Compd. 585 465
20Ashiq M NIqbal M JGul I H 2011 J. Magn. Magn. Mater. 323 259
21Choi D HAn S YLee S WShim I BKim C S 2004 Phys. Stat. Sol. 241 1736
22Chen DLiu Y LLi Y XYang KZhang H W 2013 J. Magn. Magn. Mater. 337�?38 65
23Adelskold V1938Arkiv Kemi Miner. Geol.A121
24Trukhanov A VTurchenko V OBobrikov I ATrukhanov S VKazakevich I SBalagurov A M 2015 J. Magn. Magn. Mater. 393 253
25Rai B KMishra S RNguyen V VLiu J P 2013 J. Alloys Compd. 550 198
26Tokunaga YKaneko YOkuyama DIshiwata SArima TWakimoto SKakurai KTaguchi YTokura Y2010Phys. Rev. lett.1052572
27Li WQiao XLi MLiu TPeng H 2013 Mater. Res. Bull. 48 4449
28Shams M HRozatian A S HYousefi M HValíček JŠepelák V 2016 J. Magn. Magn. Mater. 399 10
29Shannon R DPrewitt C TCrystallogr A 1969 Sect. B: Struct. Sci. 25 925
30Katlakunta SMeena S SSrinath SBououdina MSandhya RPraveena K 2015 Mater. Res. Bull. 63 58
31Kelley P LKamzin A SRomachevsky K EHue D T MChinh H DSrikanth HPhan M H 2015 J. Alloys Compd. 636 323
32Evans B JGrandjean FLilot A PVogel R HGrard A 1987 J. Magn. Magn. Mater. 67 123
33Araújo J H.Soares J MGinani M FMachado F L ACunha J B M 2013 J. Magn. Magn. Mater. 343 203
34Naranga S BSingh CBai YHudiara I S 2008 Mater. Chem. Phys. 111 225
35Li Z WYang ZWei F LZhou X ZZhao J HMorrish A H 2000 Phys. Rev. 62 6530
36Han M GOu YChen W BDeng L J 2009 J. Alloys Compd. 474 18
37Haberey FVelicescu Met al.1974Proc. Third European Conference on Hard Magnetic MaterialsSeptember 17–19, 1974Amsterdam, Netherlands70
38Ghasemi AMorisako A 2008 J. Alloys Compd. 456 485
39Xu YYang G LChu D PRuzhai H 1990 Phys. Stat. Sol. (b) 157 685
40Gossinger R 1981 Phys. Stat. Sol. (a) 66 665