Next Article in Journal
The Role of Poultry Litter and Its Biochar on Soil Fertility and Jatropha curcas L. Growth on Sandy-Loam Soil
Next Article in Special Issue
Development of Multi-Cation-Doped M-Type Hexaferrite Permanent Magnets
Previous Article in Journal
Chemical Composition and Antioxidant Activity of Asteraceae Family Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of the Magnetic Properties in Si4+-Li+-Substituted M-Type Hexaferrites for Permanent Magnets

1
Department of Materials Science & Engineering, Korea National University of Transportation, Chungju 27469, Republic of Korea
2
Department of Material Science and Engineering, Research Institute of Advanced Materials (RIAM), Seoul National University, Seoul 08826, Republic of Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(23), 12295; https://doi.org/10.3390/app122312295
Submission received: 21 October 2022 / Revised: 21 November 2022 / Accepted: 29 November 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Magnetic Materials: Characterization and Sensing Application)

Abstract

:
A series of charge-balanced Si4+-M1+,2+ (M = Mg2+, K+, Li+) substitution compositions with the chemical formulae of SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr1−yFe12−ySiyKyO19 (y = 0.05, 0.1, 0.2), and SrFe12−z(Si0.6Li0.6)zO19 (z = 0.05, 0.1, 0.2, 0.4, 0.6) were prepared using conventional ceramic processes. While the sole doping of Si with x = 0.1 (SrFe12−xSixO19 (x = 0.1)) causes a noticeable Fe2O3 phase formation, the co-doping of Si-Mg and Si-Li allow a single M-type phase formation with x up to x = 0.1 and z = 0.4, respectively. Notably, a 2.6% increase in the saturation magnetization was obtained for Si-Li-substituted SrM with z = 0.05—that is, SrFe11.95Si0.03Li0.03O19. Enhancement of the magnet performance can also be achieved when anisotropic permanent magnets are fabricated based on the substitution composition SrFe11.95Si0.03Li0.03O19. The remnant magnetic flux density was improved by 2.5% compared to that of the unsubstituted SrM (from 4207 to 4314 G). The maximum energy product (BHmax) value also increased from 4.24 to 4.46 M·G·Oe. The enhancement in permanent magnet performance is attributed to the increase in the MS of SrM by the optimal Si-Li substitution. This is a promising result because enhanced permanent magnet performance is achieved with intrinsic magnetic property improvement.

1. Introduction

Hexaferrites, a type of ferrimagnetic ceramic that emerged in the early 1950s, have attracted significant research interest in the past several decades owing to their diverse applicability in permanent magnets, magnetic recording media, and high-frequency soft magnetic devices [1,2,3]. Hexaferrite has six different crystal phases, M, U, W, X, Y, and Z, depending on the stacking structure and constituent elements. Among them, M-type hexaferrite with the basic formula (Ba,Sr)Fe12O19 has excellent phase stability, strong c-axis magnetic anisotropy, and sufficient saturation magnetization (MS); thus, it is the most utilized material in permanent magnets.
Ferrite magnets with the M-hexaferrite phase are widely used in automobiles, home appliances, and various industrial equipment around the world, which makes improving the magnet performance crucial. Several studies have been conducted to enhance the intrinsic magnetic properties, such as MS and magnetocrystalline anisotropy, by substituting various cations for SrFe12O19 (SrM) [4,5,6,7,8,9,10,11,12,13,14,15]. In the unit cell of the SrM, 24 Fe3+ ions have five different crystallographic sites—that is, one tetrahedral (4f1), three octahedral (12k, 2a, 4f2), and one hexahedral (2b) site. The electron spins of Fe3+ ions in the sites are ferrimagnetically coupled through superexchange interactions with the oxygen ions causing the spins in the 2a, 12k, and 2b sites to align in parallel to the crystallographic c axis and those in 4f1 and 4f2 sites to align anti-parallel [3].
It has been reported that La-Zn [4,5] or Mn-Zn [6] substitution into SrM increases the MS because the non-magnetic Zn2+ ions selectively replace Fe3+ ions with spin directions that are antiparallel to the net magnetization direction. It has also been proven that La-Ca-Co co-substitutions into SrM are one of most successful approaches leading to significant improvement in the uniaxial magnetocrystalline anisotropy constant (Ku) without decreasing the MS. Thus, La-Ca-Co-substituted SrM has been commercialized into high-grade magnet products [7,8,9]. However, the doping of cobalt, a highly expensive element, significantly drives up its price. Therefore, reducing material costs without compromising the magnetic properties is an important concern in the research and development of permanent magnets.
Recently, improved hard magnetic properties have been reported in hexaferrite nanomagnets prepared by sol–gel-based methods [16,17,18]. In these studies, a large coercivity (HC) > 6 kOe was achieved owing to the nanocrystallization of hexaferrire grains. It is well-known that the coercivity of a single grain (hC) increases as its size decreases until it reaches the upper limit of the critical grain diameter for a single domain (dcri). This exhibits a close reciprocal dependence of on grain size (d) as expressed in the below Equation [19].
hC = dcri/d1.08
However, for the real application of sintered magnets, a process capable of mass production through a solid-state reaction method is required rather than a chemical synthesis, such as sol–gel.
The successful development of a ferrite magnet requires not only improvement of the intrinsic hard-magnetic parameters (such as MS and Ku) but also successful process control. The improvement of intrinsic magnetic properties can be achieved through the proper cation substitution. In process control, grain growth suppression and densification must be simultaneously performed during the sintering process. SiO2 is a sintering additive that plays a crucial role in suppressing grain growth. However, if Si diffuses into the SrM matrix during the high-temperature sintering process, the MS could be lowered owing to the weakening of ferromagnetism.
In our previous study [20], M-hexaferrite sintered bodies were manufactured using two different methods by changing the order of SiO2 addition. In the first method (pre-Si process), 1 wt% of SiO2 powder was added during the mixing of the initial raw materials, and in the second method (post-Si process), the same amount of SiO2 powder was added as a sintering additive during the ball milling step after calcination. When the samples were sintered at T = 1250 °C, the MS values decreased by 4.7% and 4.8% in the post-Si and pre-Si samples, respectively, compared to that of sample where Si was not added (non-Si). The coercivity (HC) of both the pre-Si and post-Si processed samples significantly improved compared to the non-Si case.
The results of previous studies suggested that Si diffusion into the SrM matrix causes a decrease in MS, and grain growth can be effectively controlled during sintering regardless of whether SiO2 is added as a sintering additive (post-Si) or in the initial precursor mixing stage (pre-Si) [20]. In general, apart from SiO2, sintering additives, such as SrCO3 (or SrO) and CaCO3 (or CaO), are also mixed in an appropriate ratio. The simultaneous additions of CaCO3 and SrCO3 are known to be beneficial in tailoring a dense microstructure with relatively fine microstructure [15]. The SiO2 effectively suppresses grain growth and increases the HC and the Ca or Sr oxides (or carbonates) promote the densification, which increases the residual magnetic flux density (Br) of the magnet.
Therefore, Si was used as a substitute rather than as a sintering additive in this study. For satisfying the charge neutrality conditions, Si4+ and either Li+, K+, or Mg2+ ions were co-substituted into the SrM structure to find an optimal substitution composition with excellent hard magnetic properties. Thus, nominal compositions of SrFe12−2xSixMgxO19, Sr1−yFe12−ySiyKyO19, and SrFe12−z(Si0.6Li0.6)zO19 were designed.
Notably, in the composition of SrFe12−z(Si0.6Li0.6)zO19, the substitution ions are expected to be partly employed in the interstitial sites because the total number of Li+ and Si4+ ions is 20% more than that of the substitution site (Fe3+). For the sintered samples, changes in the microstructure and magnetic properties according to each substitution level were studied. For a substitution composition that exhibits optimal hard magnetic properties, anisotropic magnets were manufactured, and the performance of the permanent magnets was evaluated.

2. Materials and Methods

M-type hexaferrites with nominal compositions of SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr1−yFe12−ySiyKyO19 (y = 0.05, 0.1, 0.2), and SrFe12−z(Si0.6Li0.6)zO19 (z = 0.05, 0.1, 0.2, 0.4, 0.6) were prepared using solid-state reaction routes. The precursor powders of 99.5% Fe2O3 and 99.0% SiO2 (industrial use), 99.9% SrCO3, 99.9% Li2CO3, 99.9% MgO, and 99.9% K2CO3 (Kojundo Chemical Lab. Co., Ltd., Tokyo, Japan) were weighed according to the chemical formulae. The powder mixture was ball-milled in water for 24 h using a polypropylene jar (77 mm in diameter and 158 mm in height) and yttria-stabilized ZrO2 balls with diameters of 3, 5, and 10 mm.
The mixing ratios of the ZrO2 balls with diameters of 3, 5, and 10 mm were approximately 50, 35, and 15 wt%, and their total weight was five times that of the total sample powder. The dried powder was placed in an alumina crucible and calcined at a temperature of 1100 °C in air for 4 h. The calcined powders were crushed, sieved, ball-milled again for 20 h, and dried. For samples with a selected composition, sintering additives, such as SrCO3, CaCO3, and SiO2, were added in the second ball milling process before sintering.
The milled hexaferrite powders were pressed in a mold with a diameter of 15 mm at a pressure of 16.7 MPa to produce disk-shaped green compact pellets. The pelletized samples were sintered in air in the temperature range of 1230–1250 °C for 2 h. During the sintering process, the heating rate was 5 °C/min, and the samples were then furnace-cooled to room temperature. For the selected composition of the samples, an anisotropic sintered magnet with a diameter of approximately 35 mm was fabricated at a ferrite magnet manufacturing site (Union Materials Corporation in Korea). Here, all process conditions except for the magnetic field molding process were performed in the same manner as for preparing an isotropic sintered body. During the molding process for anisotropic magnets, a magnetic field of 10 kOe was applied for the alignment of the hexaferrite particles, and the molding pressure was 3.92 MPa.
The sample density was calculated based on the weight and geometric dimensions of the disk-shaped samples. X-ray diffraction (XRD, D8 Advance, Bruker) was performed for phase identification using a Cu Kα radiation source (λ = 0.15406 nm), and Rietveld refinement was performed on the selected samples. The fitting of the XRD patterns was conducted using the software TOPAS (Bruker, version 4.2). Field-emission scanning electron microscopy (FE-SEM, JSM-7610F, JEOL) was performed for phase identification.
The average grain size was evaluated using a random intercept method applicable to the SEM micrographs. The magnetization curves were measured using a vibrating-sample magnetometer (VSM, Lakeshore 7410) at room temperature with a sweeping magnetic field within ± 25 kOe for sintered samples with randomly oriented crystalline grains. The performance of the anisotropic magnets fabricated in a commercial magnet manufacturer (Union Materials Corp., Pohang, Republic of Korea) was characterized using a B-H loop tracer (BH-5501, Denshijiki Industry) at room temperature.

3. Results and Discussion

3.1. Crystalline Structure Analysis

In Figure 1a–d, we show the XRD patterns of the SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr1−yFe12−ySiyKyO19 (y = 0, 0.05, 0.1, 0.2), SrFe12−z(Si0.6Li0.6)z (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6), and SrFe12−xSixO19 (x = 0, 0.1) samples sintered at 1250 °C. The formation of a secondary phase of Fe2O3 is common when the substitution level increases to a certain extent. No other secondary phases were identified.
All the diffraction peaks for SrM samples were indexed based on hexagonal magnetoplumbite crystal structures with the space group P63/mmc (International Center for Diffraction Data (ICDD), Powder Diffraction File search no. 00-33-1340). When Si-Mg was co-substituted at the Fe site of the SrM (SrFe12−2xSixMgxO19), no secondary phase was formed at x = 0.05 and x = 0.1 (Figure 1a). When Si was substituted solely at the Fe site (SrFe12−xSixO19) with a substitution level of x = 0.1, the Fe2O3 phase was formed in the sample (Figure 1d).
When Si-K was co-substituted at Fe and Sr sites (Sr1−yFe12−ySiyKyO19), a single M-type phase could not be formed even at a small substitution level of y = 0.05 because K+ could not be placed at the Sr site in SrM. Thus, the effective molar ratio, [Fe + Si]/[Sr] exceeded 12, and the Fe not included in the SrM phase formed Fe2O3. Interestingly, when Li-Si was co-substituted at the Fe site (SrFe12−z(Si0.6Li0.6)zO19), a single SrM phase could be formed without the secondary phase of Fe2O3 up to z = 0.4, implying that 0.24 Si4+ and 0.24 Li+ ions are soluble in the SrM phase with a composition of SrFe11.6(Si0.24Li0.24)O19.
Therefore, it was reasonably inferred that the extra Si4+ or Li+ that could not occupy the Fe site (~17% of total substituted ions), existed in the interstitial site of the SrM lattice. Rietveld analysis results on the samples of z = 0, 0.05, and 0.4, are presented in Figure 2a–c. A single M-type phase was confirmed even for the z = 0.4 sample. The best fitting results were obtained when it was assumed that Li+ is replaced with 4f1, 4f2 sites randomly and that Si4+ is replaced with 2a sites of Fe3+. However, in this study, some substitution elements (~17%) were assumed to enter the interstitial site, which was not supported by the calculation software.
For the series of the SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr1−yFe12−ySiyKyO19 (y = 0, 0.05, 0.1, 0.2), and SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6) samples, the calculated lattice parameters a and c, their ratios c/a, and cell volumes are listed in Table 1. The a and c values were calculated from the dhkl value corresponding to the (107) and (114) peaks according to the following equation:
d h k l = { 4 ( h 2 + h k + k 2 ) 3 a 2 + l 2 c 2 } 1 / 2 .
where dhkl is the interplanar spacing, and h, k, and l are the Miller indices. The changes in these parameters according to the substitution levels, x, y, and z, are shown in Figure 3. For the three cases (Figure 3a–c), as the substitution amount of Si increased, the value of a decreased, and the c value and c/a ratio increased gradually. The cell volume tended to decrease with increasing substitution levels.
In the case of the Si-Mg substitution, the cell volume decreased by approximately 0.35% at x = 0.2. However, in the case of K-Si and Li-Si substitutions, the cell volume did not change significantly. It was 0.1% or lower at the same Si doping level.

3.2. Microstucture Analysis

The changes in the microstructure according to the substitution elements and their amounts are shown in Figure 4a–l. The average grain size is shown in the lower-right corner of each figure. Most samples exhibited a porous microstructure because they were sintered without sintering additives. In Figure 4a–d, the microstructure of the Si-Mg-substituted SrM is shown with increasing substitution levels (x). When x = 0.05 and 0.1, the average grain size (Dave) decreased to 0.78 and 0.65 μm, respectively. It increased to Dave = 0.98 μm at x = 0.2. The microstructures of the Si-K-substituted SrM samples are shown in Figure 4e–g, and those of Si-Li are shown in Figure 4h–l. In the case of Si-K doping, the grain size was finely controlled when the substitution amount was y ≤ 1, whereas in the case of Si-Li substitution, a fine microstructure could be obtained when the substitution amount was z ≥ 0.2.

3.3. Magnetic Properties and Magnet Performances

Figure 5 shows the magnetic hysteresis curves of the SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) and SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6) samples sintered at 1250 °C. In the case of the Si-K-substituted SrM samples, a considerable amount of Fe2O3 was formed as a secondary phase even with a small doping level z (Figure 1b; therefore, the magnetic properties were not evaluated. The MS and HC values with respect to x and z are plotted in Figure 5c,d. The sintered sample densities as well as MS and HC values of the samples are listed in Table 2.
In the case of Si-Mg substitution, the MS slightly increased at x = 0.05 and HC increased significantly. Then, MS and HC tended to decrease as x further increased to 0.2. As the spin moment contributing to the magnetization value in SrM originated from the Fe3+ ions, the MS decreased when the Fe content decreased owing to a further increase in the Si-Li substitution. The most notable improvement in the magnetic properties was observed in the Si-Li-substituted sample with z = 0.05. In this sample, the MS value increased significantly from 72.08 to 73.92 emu/g (2.6%), and the HC value also increased slightly.
This improvement in MS is unusual because when Si penetrates the M-type hexaferrite lattice, it is expected to decrease the MS [20].The MS value is an intrinsic magnetic property that depends on the composition and lattice characteristics, unlike HC, which is greatly affected by the microstructure. Here, the MS values (in emu/g unit) were obtained from samples that had not been densified. A sintered density that is 95% or higher than the theoretical density of approximately 5.1 g/cm3 [17] is required to become a permanent magnet with excellent Br.
Therefore, in the next experiment, isotropic permanent magnets using various sintering additives were fabricated with the optimized composition of SrFe11.95Si0.03Li0.03O19, which exhibited the highest MS value with a relatively high HC. Figure 6a shows the demagnetization curves of the hexaferrite samples sintered with various additives, including 0.5 wt% SrCO3, 0.75 wt% SrCO3, 1 wt% SrCO3, 0.5 wt% SrCO3 + 0.5 wt% CaCO3, 1 wt% SrCO3 + 0.5 wt% SiO2, 1 wt% SrCO3 + 0.5 wt% Co3O4, and 1 wt% CaCO3 + 0.5 wt% SiO2, respectively. After measuring the full hysteresis in the range of H = 20 kOe, the demagnetization curves in the second quadrant were obtained. The density (ρ), remanent magnetization (Br), and coercivity (HC) of the sintered hexaferrite samples are presented in Table 3.
The sintering density (ρ = 4.7 ~ 5.0 g/cm3, Table 3) significantly increased when the sintering additives were used compared to when they were not (ρ = 2.7~3.9g/cm3, Table 2). Notably, the sintering temperature was set to 1230 °C (20 °C lower than the case in Table 2). We also found that a sufficient HC value could not be obtained without adding 0.5 wt% SiO2. In addition, a sample composition of La-Co-Si-Li-substituted SrM, Sr0.8La0.2Fe11.75Co0.2Si0.03Li0.03O19 was synthesized using the same procedure and sintered with two set of additives: 1 wt% CaCO3 + 0.5 wt% SiO2 and 1 wt% SrCO3 + 0.5 wt% SiO2.
Since it is well-known that La-Co substitution into SrM enhances the hard magnetic properties [7,8,9], the composition Sr0.8La0.2Fe11.75Co0.2Si0.03Li0.03O19 was specifically designed for producing high-performance permanent magnets with the expectation of combining the effects of La-Co and Si-Li substitution. The demagnetization curves of La-Co-Si-Li-substituted SrM are presented in Figure 6b, and the magnetic parameters are listed in Table 3. The La-Co-Si-Li-added SrM samples showed significantly improved HC values compared with the Si-Li-substituted SrM. This is because La-Co substitution increases the Ku of SrM.
Finally, anisotropic sintered magnets with the compositions SrFe12O19, SrFe11.95Li0.03Si0.03O19, and Sr0.8La0.2Fe11.75Co0.2Si0.03Li0.03O19 were fabricated by adding a magnetic field pressing process. Sintering additives 0.5wt% SiO2 + 1.0wt% CaCO3 are commonly used. The demagnetization curves (4πM-H and B-H) of the anisotropic magnets are shown in Figure 7, and the magnet parameters are presented in Table 4. The Br of Si-Li-substituted SrM (4314 G) increased by 2.5% compared to that of unsubstituted SrM (4207 G), and the HC also slightly increased. Thus, the maximum energy product (BHmax) value increased from 4.24 to 4.46 M·G·Oe.
The 2.5% increase in Br can be attributed to the increase in the MS of the Si-Li-substituted SrM (z = 0.05) sample, as shown in Figure 3d. La-Co-Si-Li-substituted SrM exhibited a high HC value; however, the bHC and BHmax values were significantly lower than those of Si-Li SrM due to the poor squareness of the demagnetization curve. The squareness of the demagnetization curves is closely related to the grain orientation, which is determined during the magnetic field pressing or sintering processes. Therefore, the performance of the magnet (Br and BHmax) could be improved by further process optimization.

4. Conclusions

In this study, M-type hexaferrites with the chemical formulae SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr1−yFe12−ySiyKyO19 (y = 0.05, 0.1, 0.2), and SrFe12−z(Si0.6Li0.6)zO19 (z = 0.05, 0.1, 0.2, 0.4, 0.6) were prepared using conventional solid-state reaction processes. A single M-hexaferrite phase was formed in the charge-balanced Si4+-Mg2+ and Si4+-Li+ co-substitution for Fe3+ up to x = 0.1 and z = 0.4, respectively. A fine microstructure with Dave of less than 0.8 μm was obtained when the substitution level was in the range of x ≤ 0.1 in the Si-Mg-substituted SrM, whereas, in the case of Si-Li substitution, it could be obtained at z ≥ 0.2.
When the magnetic properties according to x and z were evaluated through M-H measurements, the best hard magnetic properties were obtained at z = 0.5 with the MS value increased by 2.6%. When an anisotropic permanent magnet was prepared and evaluated for the optimal composition SrFe11.5Si0.03Li0.03O19, Br improved by 2.5% from 4207 to 4314 G compared to the unsubstituted SrM. The maximum energy product (BHmax) value also increased from 4.24 to 4.46 M·G·Oe. The enhancement of the intrinsic magnetic property of SrM by the substitution of Li-Si, a new composition separate from La-Co, is a promising result and is expected to see further improvement through processing optimization.

Author Contributions

Conceptualization, Y.-M.K.; Funding acquisition, Y.-M.K.; Investigation, Y.-M.K. and S.-I.Y.; Methodology, J.-Y.Y., K.-H.L., and Min-Ho Kim; Project administration, Y.-M.K. and Min-Ho Kim; Software, J.-Y.Y. and K.-H.L.; Supervision, Y.-M.K. and S.-I.Y.; Writing—original draft, J.-Y.Y. and Y.-M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Basic Science Research Capacity Enhancement Project (National Research Facilities and Equipment Center) through the Korea Basic Science Institute funded by the Ministry of Education (Grant No. 2019R1A6C1010047) and by the Korea Basic Science Institute (National Research Facilities and Equipment Center) grant funded by the Ministry of Education (grant No. 2021R1A6C103A367).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to Min-Ho Kim in the R&D Team, Union Materials Corp. for his support of the anisotropic ferrite magnet fabrication.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Went, J.J.; Rathenau, G.W.; Gorter, E.W.; Van Oosterhout, G.W. Ferroxdure. A Class of Permanent Magnetic Materials. Philips Tech. Rev. 1952, 13, 194–208. [Google Scholar]
  2. Ashiq, M.N.; Iqbal, M.J.; Najam-ul-Haq, M.; Gomez, P.H.; Qureshi, A.M. Synthesis, magnetic and dielectric properties of Er–Ni doped Sr-hexaferrite nanomaterials for applications in High density recording media and microwave devices. J. Magn. Magn. Mater. 2012, 324, 15–19. [Google Scholar] [CrossRef]
  3. Pullar, R.C. Hexagonal ferrites: A review of the synthesis, properties and applications of hexaferrite ceramics. Prog. Mater. Sci. 2012, 57, 1191–1334. [Google Scholar] [CrossRef]
  4. Taguchi, H. Recent improvements of ferrite magnets. J. Phys. IV Fr. 1997, 7, C1-299–C1-302. [Google Scholar] [CrossRef]
  5. Bai, J.; Liu, X.; Xie, T.; Wei, F.; Yang, Z. The effects of La–Zn substitution on the magnetic properties of Sr-magnetoplumbite ferrite nano-particles. Mater. Sci. Eng. 2000, 68, 182–185. [Google Scholar] [CrossRef]
  6. Kang, Y.-M.; Kwon, Y.-H.; Kim, M.-H.; Lee, D.-Y. Enhancement of magnetic properties in Mn-Zn substituted M-type Sr-hexaferrites. J. Magn. Magn. Mater. 2015, 382, 10–14. [Google Scholar] [CrossRef]
  7. Kools, F.; Morel, A.; Grössinger, R.; Le Breton, J.M.; Tenaud, P. LaCo-substituted ferrite magnets, a new class of high-grade ceramic magnets; intrinsic and microstructural aspects. J. Magn. Magn. Mater. 2002, 242–245, 1270–1276. [Google Scholar] [CrossRef]
  8. Ogata, Y.; Takami, T.; Kubota, Y. Development of La-Co substituted ferrite magnets. J. Jpn. Soc. Powder Metall. 2003, 50, 636–641. [Google Scholar] [CrossRef] [Green Version]
  9. Kobayashi, Y.; Hosokawa, S.; Oda, E.; Toyota, S. Magnetic properties and composition of Ca-La-Co M-type ferrites. J. Jpn. Soc. Powder Metall. 2008, 55, 541–546. [Google Scholar] [CrossRef] [Green Version]
  10. Kang, Y.-M.; Moon, K.-S. Magnetic properties of Ce-Mn substituted M-type Sr-hexaferrites. Ceram. Inter. 2015, 41, 12828–12834. [Google Scholar] [CrossRef]
  11. Barrera, V.; Betancourt, I. M-type hexaferrites with enhanced coercivity. IEEE Trans. Magn. 2013, 49, 4630–4633. [Google Scholar] [CrossRef]
  12. Wang, H.Z.; Yao, B.; Xu, Y.; He, Q.; Wen, G.H.; Long, S.W.; Fan, J.; Li, G.D.; Shan, L.; Lie, B.; et al. Improvement of the coercivity of strontium hexaferrite induced by substitution of Al3+ ions for Fe3+ ions. J. Alloys Comp. 2012, 537, 43–49. [Google Scholar] [CrossRef]
  13. Ahn, K.; Ryu, B.; Korolev, D.; Kang, Y.J. Substantial enhancement in intrinsic coercivity on M-type strontium hexaferrite through the increase in magneto-crystalline anisotropy by co-doping of group-V and alkali elements. Appl. Phys. Lett. 2013, 103, 242417. [Google Scholar] [CrossRef]
  14. Kang, Y.-M. High saturation magnetization in La–Ce–Zn–doped M-type Sr-hexaferrites. Ceram. Int. 2015, 41, 4354–4359. [Google Scholar] [CrossRef]
  15. Moon, K.-S.; Kang, Y.-M. Structural and magnetic properties of Ca-Mn-Zn-substituted M-type Sr-hexaferrites. J. Eur. Ceram. Soc. 2016, 36, 3383–3389. [Google Scholar] [CrossRef]
  16. Shekhawat, D.; Singh, A.K.; Roy, P.K. Structural and electro-magnetic properties of high (BH)max La-Sm substituted Sr-hexaferrite for brushless DC electric motors application. J. Mol. Struct. 2019, 1179, 787–794. [Google Scholar] [CrossRef]
  17. Eikeland, A.Z.; Stingaciu, M.; Mamakhel, A.H.; Saura-Múzquiz, M.; Christensen, M. Enhancement of magnetic properties through morphology control of SrFe12O19 nanocrystallites. Sci. Rep. 2018, 8, 7325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Lee, J.; Lee, E.J.; Hwang, T.-Y.; Kim, J.; Choa, Y.-H. Anisotropic characteristics and improved magnetic performance of Ca–La–Co-substituted strontium hexaferrite nanomagnets. Sci. Rep. 2020, 10, 15929. [Google Scholar] [CrossRef] [PubMed]
  19. Nishio, H.; Minachi, Y.; Yamamoto, H. Effect of factors on coercivity in Sr-La-Co sintered ferrite magnets. IEEE Trans. Magn. 2009, 45, 5281–5288. [Google Scholar] [CrossRef]
  20. Moon, K.-S.; Yu, P.-Y.; Kang, Y.-M. Microstructure and magnetic properties of La-Ca-Co substituted M-type Sr-hexaferrites with controlled Si diffusion. Appl. Sci. 2020, 10, 7570. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of (a) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), (b) Sr1−yFe12−ySiyKyO19 (y = 0, 0.05, 0.1, 0.2), (c) SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6), and (d) SrFe12−xSixO19 (x = 0, 0.1) samples sintered at 1250 °C.
Figure 1. X-ray diffraction patterns of (a) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), (b) Sr1−yFe12−ySiyKyO19 (y = 0, 0.05, 0.1, 0.2), (c) SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6), and (d) SrFe12−xSixO19 (x = 0, 0.1) samples sintered at 1250 °C.
Applsci 12 12295 g001
Figure 2. (ac) XRD patterns with Rietveld refinement for SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.4) samples.
Figure 2. (ac) XRD patterns with Rietveld refinement for SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.4) samples.
Applsci 12 12295 g002
Figure 3. Change (%) in the lattice parameters a, c, and cell volume of (a) SrFe12−2xSixMgx, (b) Sr1−yFe12−ySiyKy, and (c) SrFe12−z(Li0.6Si0.6)z samples sintered at 1250 °C.
Figure 3. Change (%) in the lattice parameters a, c, and cell volume of (a) SrFe12−2xSixMgx, (b) Sr1−yFe12−ySiyKy, and (c) SrFe12−z(Li0.6Si0.6)z samples sintered at 1250 °C.
Applsci 12 12295 g003
Figure 4. SEM micrographs of the (ad) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), (a,eg) Sr1−yFe12−ySiyKyO19 (y = 0, 0.05, 0.1, 0.2), and (a,hl) SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6) samples sintered at 1250 °C.
Figure 4. SEM micrographs of the (ad) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), (a,eg) Sr1−yFe12−ySiyKyO19 (y = 0, 0.05, 0.1, 0.2), and (a,hl) SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6) samples sintered at 1250 °C.
Applsci 12 12295 g004
Figure 5. M-H curves of (a) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) and (b) SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6) samples sintered at 1250 °C, and (c,d) plots of MS and HC vs. x and vs. z. Here, the MS value is defined as the magnetization (M) value at H = 2.5 kOe.
Figure 5. M-H curves of (a) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) and (b) SrFe12−z(Si0.6Li0.6)zO19 (z = 0, 0.05, 0.1, 0.2, 0.4, 0.6) samples sintered at 1250 °C, and (c,d) plots of MS and HC vs. x and vs. z. Here, the MS value is defined as the magnetization (M) value at H = 2.5 kOe.
Applsci 12 12295 g005
Figure 6. Demagnetization curves (4πM-H) of (a) SrFe11.95Si0.03Li0.03O19 and (b) Sr0.8La0.2Fe11.75Co0.2Si0.03Li0.03O19 samples sintered at 1230 °C with various sintering additives.
Figure 6. Demagnetization curves (4πM-H) of (a) SrFe11.95Si0.03Li0.03O19 and (b) Sr0.8La0.2Fe11.75Co0.2Si0.03Li0.03O19 samples sintered at 1230 °C with various sintering additives.
Applsci 12 12295 g006
Figure 7. Demagnetization curves (4πM-H and B-H) of the anisotropic magnets, SrFe11.95Si0.03Li0.03O19 and Sr0.8La0.2Fe11.75Co0.2Si0.03Li0.03O19, sintered at 1230 °C.
Figure 7. Demagnetization curves (4πM-H and B-H) of the anisotropic magnets, SrFe11.95Si0.03Li0.03O19 and Sr0.8La0.2Fe11.75Co0.2Si0.03Li0.03O19, sintered at 1230 °C.
Applsci 12 12295 g007
Table 1. The lattice parameters a, c, c/a ratios, and lattice volumes (vol.) of the SrFe12−2xSixMgx, Sr1−yFe12−ySiyKy, and SrFe12−z(LizSiz)0.6 samples sintered at 1250 °C.
Table 1. The lattice parameters a, c, c/a ratios, and lattice volumes (vol.) of the SrFe12−2xSixMgx, Sr1−yFe12−ySiyKy, and SrFe12−z(LizSiz)0.6 samples sintered at 1250 °C.
Compositionx, y, za (Å) c (Å) c/a vol. (Å3)
SrFe12−2xSixMgxO1905.88523.0553.917691.50
0.055.87923.0593.922690.21
0.15.88123.0723.923691.06
0.25.87023.0923.934689.02
Sr1−yFe12−ySiyKyO190.055.88323.0793.923691.74
0.15.88223.0663.921691.17
0.25.88023.0763.925690.90
SrFe12−z(Si0.6Li0.6)zO190.055.88423.0563.919691.20
0.15.88223.0663.921691.19
0.25.88123.0753.924691.09
0.45.87823.0763.926690.53
0.65.88023.0833.926691.18
Table 2. The sintered sample density (ρ), saturation magnetization (Ms), and coercivity (HC) of hexaferrite samples, SrFe12−2xSixMgx and SrFe12−z(LizSiz)0.6 sintered at 1250 °C. The error bar for MS is ± 0.2% and for HC is ±1% of the original values.
Table 2. The sintered sample density (ρ), saturation magnetization (Ms), and coercivity (HC) of hexaferrite samples, SrFe12−2xSixMgx and SrFe12−z(LizSiz)0.6 sintered at 1250 °C. The error bar for MS is ± 0.2% and for HC is ±1% of the original values.
Compositionx, zρ (g/cm3)MS (emu/g)HC (Oe)
SrFe12−2xSixMgxO1902.7572.083562
0.052.8772.153912
0.12.9170.463402
0.24.0766.492986
SrFe12−z(LizSiz)0.6O1902.7572.083559
0.053.0973.923713
0.13.2472.363393
0.23.2370.53904
0.43.9070.73809
0.63.8464.153216
Table 3. The sintered sample density (ρ), remanent magnetization (Br), and coercivity (HC) of hexaferrite samples SrFe11.95Li0.03Si0.03O19 and Sr0.8La0.2Fe11.75Co0.2Li0.03Si0.03O19 with different additives sintered at 1230 °C.
Table 3. The sintered sample density (ρ), remanent magnetization (Br), and coercivity (HC) of hexaferrite samples SrFe11.95Li0.03Si0.03O19 and Sr0.8La0.2Fe11.75Co0.2Li0.03Si0.03O19 with different additives sintered at 1230 °C.
CompositionAdditivesρ (g/cm3)Br (G)HC (Oe)
SrFe11.95Li0.03Si0.03O19SrCO3 0.5wt%4.9226292878
SrCO3 0.75 wt%4.9326312554
SrCO3 1 wt%5.0424591810
SrCO3 0.5 wt% + CaCO3 0.5 wt%5.0425962129
SrCO3 1 wt% + SiO2 0.5 wt%4.7025094149
SrCO3 1 wt% + Co3O4 0.5 wt%5.0423751914
CaCO3 1 wt% + SiO2 0.5 wt%4.8025284165
Sr0.8La0.2Fe11.75Co0.2Li0.03Si0.03O19SiO2 0.5 wt% + SrCO3 1.0 wt%4.6824204786
SiO2 0.5 wt% + CaCO3 1.0 wt%4.9225824608
Table 4. The magnet density, intrinsic coercivity (iHC), coercivity in BH (bHC), remanence magnetic flux density (Br), and maximum energy product (BHmax) of the sintered anisotropic magnets.
Table 4. The magnet density, intrinsic coercivity (iHC), coercivity in BH (bHC), remanence magnetic flux density (Br), and maximum energy product (BHmax) of the sintered anisotropic magnets.
Magnet Compositionρ
(g/cm3)
iHC
(Oe)
bHC
(Oe)
Br
(G)
BHmax
(M·G·Oe)
SrFe12O194.953182300242074.24
SrFe11.95Li0.03Si0.03O194.943206300943144.46
Sr0.8La0.2Fe11.75Co0.2Li0.03Si0.03O195.023897313842933.98
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

You, J.-Y.; Lee, K.-H.; Kang, Y.-M.; Yoo, S.-I. Enhancement of the Magnetic Properties in Si4+-Li+-Substituted M-Type Hexaferrites for Permanent Magnets. Appl. Sci. 2022, 12, 12295. https://doi.org/10.3390/app122312295

AMA Style

You J-Y, Lee K-H, Kang Y-M, Yoo S-I. Enhancement of the Magnetic Properties in Si4+-Li+-Substituted M-Type Hexaferrites for Permanent Magnets. Applied Sciences. 2022; 12(23):12295. https://doi.org/10.3390/app122312295

Chicago/Turabian Style

You, Jin-Young, Kang-Hyuk Lee, Young-Min Kang, and Sang-Im Yoo. 2022. "Enhancement of the Magnetic Properties in Si4+-Li+-Substituted M-Type Hexaferrites for Permanent Magnets" Applied Sciences 12, no. 23: 12295. https://doi.org/10.3390/app122312295

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop