Next Article in Journal
Dynamic Pathway Selection Mechanisms of Brain Networks
Next Article in Special Issue
A Study of Magnetic Mill Productivity
Previous Article in Journal
Research on Mechanical Characteristics of Slope Reinforcement by Spatial Arc Crown Beam Composite Supporting Structure
Previous Article in Special Issue
Enhancement of the Magnetic Properties in Si4+-Li+-Substituted M-Type Hexaferrites for Permanent Magnets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of Multi-Cation-Doped M-Type Hexaferrite Permanent Magnets

Department of Materials Science & Engineering, Korea National University of Transportation, Chungju 27469, Republic of Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(1), 295; https://doi.org/10.3390/app13010295
Submission received: 30 November 2022 / Revised: 20 December 2022 / Accepted: 23 December 2022 / Published: 26 December 2022
(This article belongs to the Special Issue Magnetic Materials: Characterization and Sensing Application)

Abstract

:
We report enhanced permanent magnet performance for multi-cation–substituted M-type Sr-hexaferrites (SrM) prepared using conventional ceramic processes. The final cation composition, Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19, could be derived through stepwise and systematic cation composition designs, processing, and characterization. The hexaferrites sample sintered in the temperature range of 1200–1220 °C showed an enhanced coercivity (HC) of approximately 4.0 kOe and a residual magnetic flux density (Br) of 2.5–2.6 kG. When samples of the same composition were fabricated into anisotropic magnets through a magnetic-field molding process, performance parameters of Br = 4.42 kG, HC = 3.57 kOe, and BHmax = 4.70 M·G·Oe were achieved, a significant improvement over Br = 4.21 kG, HC = 3.18 kOe, and BHmax = 4.24 M·G·Oe for the non-substituted SrFe12O19 magnet processed under optimized conditions.

1. Introduction

M-type hexaferrites, which have the basic formula (Ba,Sr)Fe12O19 and were discovered in the early 1950s, have been utilized as permanent magnets owing to their excellent chemical stability and proper hard-magnetic properties, such as strong c-axis magnetic anisotropy and sufficient saturation magnetization (MS) [1,2]. Ferrite magnets using the M-type hexaferrite phase account for the largest amount of permanent magnet usage worldwide, which makes enhancing magnet performance very important. Many studies have been conducted to enhance the magnetic properties of M-type hexaferrites, such as saturation magnetization (MS) and the uniaxial magnetocrystalline anisotropy constant (Ku), by substituting various cations. It has been reported that the optimal substitution of La-Zn [3,4], Mn-Zn [5], La-Ce-Zn [6], La-Mn-Zn [7], and Zn-Zr [8] into M-type hexaferrites increases the MS value. One possible reason for the increase in MS is that non-magnetic Zn2+ ions selectively replace Fe3+ ions with spin directions that are antiparallel to the net magnetization direction. It has also been reported that the substitution of Ce-Mn [9], Sm [10], Nd-Nb-Zn [11], Ga-Zn [12], Nb [13], Ce-Dy [14], Co-Al [15], Al [16], and Ho [17] enhances the Ku value or coercivity (HC) of M-type hexaferrites. However, MS and Ku do not improve simultaneously in all cases. Generally, in magnetic materials, an increase in MS causes a decrease in Ku, and an increase in Ku causes a decrease in MS. It is well-known that La-Ca-Co-substituted M-type hexaferrites exhibit significantly enhanced hard-magnetic properties because their magnetocrystalline anisotropy increases without scarifying the MS [18,19,20]. Therefore, La-Ca-Co-substituted Sr-hexaferrites (SrM) have been developed into high-grade ferrite magnet products. However, no meaningful progress has been made in the research and development of commercial ferrite magnets for more than 15 years. In addition, owing to the recent surge in the price of cobalt raw materials, the development of ferrite magnets with excellent performance without the use of cobalt is strongly required in the industry.
The goal of this study was to determine the optimal composition for improving the hard-magnetic characteristics of multi-cation-substituted SrM. First, we sought to determine the optimal doping amount of Si-Mg in SrM. Normally, SiO2 is used as a sintering additive to control grain growth. However, our previous studies [21] have shown that the Si can be applied as substitution elements by co-substituting Si4+ and Mg2+ at Fe3+ sites in SrM. The co-substitution of a Si4+-Mg2+ pair for two Fe3+ ions in SrM satisfies charge neutrality. To further improve the characteristics, the La-Ca-Co substitution was combined with the optimal level of Si-Mg substitution while controlling the Fe content. Finally, we sought to determine the sintering conditions required for the replacement of Co with Mn to obtain excellent hard-magnetic properties. Based on the obtained optimal cation composition and sintering process conditions, anisotropic magnets were fabricated through magnetic-field molding and their performance as ferrite magnets was evaluated.

2. Materials and Methods

M-type hexaferrites with nominal compositions of SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) were prepared using solid-state reaction routes. Precursor powders of SrCO3, CaCO3, MgO, and Co3O4 (all 99.9% purity, Kojundo Chemical Lab. Co., Ltd.), Fe2O3 (~99.5%, industrial use), and SiO2 (99.0%, industrial use) were weighed according to their chemical formulas. Each powder mixture was placed in a polypropylene bottle with yttria-stabilized ZrO2 balls and deionized (DI) water and balls milled at a rotation speed of 100 rpm for 24 h. Then, the mixed powder was dried in an oven, placed in an alumina crucible, and calcined at a temperature of 1100 °C in air for 4 h. The calcined powders were ground, sieved, ball-milled again for 20 h, and dried. For selected samples, additional powder milling was conducted for 2–6 h with a high-speed (200 rpm) milling machine (PM100, Retsch, Haan, Germany). The ball-milled powders were pressed in a disk-shaped mold with a diameter of 15.0 mm at a pressure of approximately 15 MPa to produce green compact pellets. The pelletized samples were sintered in air in the temperature range of 1200–1230 °C for 2 h. For the selected cation composition, anisotropic sintered magnets with diameters of approximately 35 mm were fabricated at a commercial ferrite magnet manufacturing site (Union Materials Corporation, Seoul, Korea). Here, all process conditions except for the magnetic-field molding process were conducted under the same conditions as those mentioned above. During the molding process for the anisotropic magnet, a magnetic field of 10 kOe was applied to align the hexaferrite particles.
The sample density was calculated based on the weight and geometric dimensions of the disk-shaped samples. X-ray diffraction (XRD; D2 Phaser, Bruker, Mannnheim, Germany) was conducted for phase identification using a Cu-Kα radiation source (λ = 0.15406 nm). Field-emission scanning electron microscopy (FE-SEM; JSM-7610F, JEOL, Tokyo, Japan) was used for phase identification. The magnetic hysteresis curves were measured using a vibrating-sample magnetometer (VSM; Lakeshore 7410, Westerville, OH, USA) and a B-H loop tracer (BH-5501, Denshijiki Industry, Ukima, Japan) at room temperature. For the VSM and B-H loop tracer measurements, fragments of sintered samples weighing approximately 0.1 g and disk-shaped sintered pellets with diameters of approximately 10.5 and 35 mm, respectively, were used.

3. Results and Discussion

3.1. Crystalline Structure and Microstructure Analysis

Figure 1a shows the powder XRD patterns of the SrFe12–2xSixMgxO19 (x = 0.0, 0.05, 0.1, 0.2) samples sintered at 1230 °C. A secondary phase peak of Fe2O3 is observed as the substitution level increases from x = 0.0 to x = 0.1, and it becomes noticeably larger at x = 0.2. The powder XRD patterns of the Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0.0, 1.0, 1.5, and 2.0) samples are shown in Figure 1b. Fe2O3 peaks are found at y = 0, and they decrease significantly at y = 1.0; no Fe2O3 peak is found at x ≥ 1.5. The diffraction peaks for the M-type hexaferrites were indexed based on hexagonal magnetoplumbite crystal structures with space group P63/mmc [International Center for Diffraction Data (ICDD), Powder Diffraction File no. 00-33-1340]. As shown in Figure 1c, no secondary phase peaks were observed in the XRD patterns of the Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) samples. The samples without Fe2O3 peaks are considered to have a single-phase M-phase.
Figure 2 shows the magnetic hysteresis curves of the SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) samples sintered at 1230 °C measured using VSM. The MS and HC values with respect to x are shown in Figure 2b. The sintered sample densities and the MS and HC values of the sintered samples are presented in Table 1. All the samples were sintered without any sintering additives, and thus the densities of the samples were much lower than the theoretical density of SrM (~5.1 g/cm3). Nevertheless, because MS was measured using VSM in units of emu/g, the densities of the samples did not affect the value of MS. The MS value increased slightly at x = 0.05, whereas the HC value increased significantly. Then, the MS and HC values decreased as x increased to 0.2. Because HC increased without a decrease in the MS value at a Si-Mg substitution amount of x = 0.05, this value of x is considered to be a very effective substitution composition for a permanent magnet. The MS value is an intrinsic magnetic property that depends on cation substitution, and the possible mechanism of the MS increase may have been the preferred substitution of non-magnetic Si4+ or Mg2+ into the down-spin site of Fe3+ [4,5,6].
The microstructures of the SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) samples sintered at 1230 °C are presented in Figure 3a–d. The samples show fine microstructures at x = 0.05 and 0.1 compared to the x = 0 sample, and they are very porous from x = 0 to 0.1. The Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2) samples sintered at 1230 °C showed a much denser microstructure with larger grains (Figure 3e–h). Similar results showing that grain growth and densification are effectively enhanced by doping with Ca [22] or LaCaCo [23,24] have been previously reported. The sintering densities of the samples are presented in Table 2. When the Fe dependency y increased to 2.0 in the Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 samples (Figure 3h), significant grain growth occurred. From the XRD patterns shown in Figure 1b, it was verified that Fe2O3 was present at x = 0 and 0.1 but lost at x = 1.5. When a small amount of Fe2O3 is present in the M-type hexaferrite, grain growth is well-controlled. However, in the y = 2.0 sample, Fe2O3 was exhausted, and thus the Fe content would be insufficient even in the mother phase (M-phase). It is believed that the lack of Fe content promotes grain growth of SrM [23].

3.2. Magnetic Properties and Magnet Performances

The demagnetization curves (4πM vs. H) of the sintered Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2) and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) hexaferrites are presented in Figure 4a–d, and the Br and HC values are presented in Table 2. The magnetization curves (4πMH) were measured using a B-H loop tracer on the disk samples. In the B-H loop tracer measurement, the 4πM data were gathered in Gaussian units, unlike those obtained in emu/g units for the VSM measurements shown in Figure 2. The M (emu/g) values can be converted to 4πM (G) using the following equation:
4πM (G) = 4πM (emu/g) × ρ,
where ρ is the sintered sample density (g/cm3). The magnetic flux density B (G), magnetization (4πM), and applied magnetic field (H) in the cgs unit system have the following relations [25]:
B (G) = 4πM (G) + H (Oe)
The [G] and [Oe] units can be considered as the same physical quantity in the cgs unit system. Therefore, a sintering density close to the theoretical density is required for a permanent magnet to produce a high magnetic flux density. As shown in Figure 4a, the y = 0 sample exhibited a relatively low Br value of 2.37 kG owing to the presence of a non-magnetic second phase of Fe2O3, which is shown in Figure 1b. The highest Br value of 2.69 kG was obtained for the y = 1.5 sample with no second phase. The HC values of the Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 samples increased from 3.78 to 3.91 kOe at y = 1.0, and it decreased with increasing y. The sudden decrease in HC at y = 2.0 can be attributed to significant grain growth (Figure 3h). As mentioned above, the lack of Fe in the M-phase is believed to promote grain growth.
Figure 4. (ad) Demagnetization curves (4πM vs. H) of the sintered Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) hexaferrites.
Figure 4. (ad) Demagnetization curves (4πM vs. H) of the sintered Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) hexaferrites.
Applsci 13 00295 g004
Next, the magnetic properties were evaluated by substituting an amount of Co with Mn in a fixed composition of y = 1.5, in which a high Br value was obtained without a Fe2O3 secondary phase. Figure 4b,c shows the 4πM-H curves of the Sr0.4Ca0.3La0.3Fe10.2Co0.2−xMnxSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) hexaferrites sintered at 1230 °C and 1210 °C, respectively. The ρ, Br, and HC values of the samples are presented in Table 2. In Figure 4b, the Mn-doped samples (z = 0.05, 0.1, 0.2) sintered at 1230 °C exhibit poor demagnetization curves, with significantly decreased Br and HC values. However, when these samples were sintered at 1210 °C, the Br and HC values of the Mn-doped samples were greatly enhanced. The z = 1.0 sample showed a significantly improved HC value of 4150 kOe and a relatively high Br value of 2.53 kG. Here, because the value of Br is proportional to the sintering density, there is room for improvement in Br when the sintering density is increased. To improve the sintering density, additional high-energy ball-milling was performed for 2, 4, and 6 h on a sample with a composition of z = 1.0 (Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19) and the resulting 4πM-H curves were measured; the results are shown in Figure 4d. As the ball-milling time increased from 2 to 6 h, the sintering density and Br value increased. Much of the change in the demagnetization curves shown in Figure 4b,c can be understood from the microstructure observations. The SEM micrographs of the Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) samples sintered at 1230 °C and 1210 °C are presented in Figure 5a–d and Figure 5e–h, respectively. Comparing the microstructures of samples with the same composition sintered at 1210 °C and 1230 °C, it can be seen that the grains of the sample sintered at 1230 °C grew slightly more than those of the sample sintered at 1210 °C, as expected. However, this is insufficient to explain the collapse of the demagnetization curve of the Mn-doped samples (z = 0.05, 0.1, 0.2) shown in Figure 4b. The SEM micrographs with low magnetization for the z = 0.05 and 0.2 samples sintered at 1210 °C and 1230 °C are presented in Figure 5i–l. The sample sintered at 1210 °C shows normal grains throughout, whereas the sample sintered at 1230 °C shows the coexistence of a region of normal grains and a region with abnormally large grains. Therefore, the poor demagnetization curves for the Mn-doped samples Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0.05, 0.1, 0.2) sintered at 1230 °C (Figure 4b) are attributed to the presence of abnormally large grains. Abnormal grain growth (AGG) can occur when a few large grains grow faster than normal grains [22]. The occurrence of AGG depends on the sintering temperature and substitution composition [17]. A similar effect of doping level on the occurrence of AGG has been reported in hexaferrite materials, wherein abnormal grain growth occurred depending on the doping level of Ca or Si [7,22,23,24]. If the sample contains large grains owing to the occurrence of AGG, the HC value of the magnet will be greatly reduced, making it unsuitable for use as a permanent magnet. HC is greatly reduced in samples where AGG occurs because a reverse magnetic domain is easily generated within a large grain, and demagnetization occurs quickly owing to the easy movement of the magnetic domain wall.
Based on the experimental results, the cation substitution composition and sintering temperature were optimized. The Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 (x = 0.05, y = 1.5, z = 1.0) samples sintered at 1210 °C exhibited excellent hard-magnetic properties. To secure process window for sintering temperature, the microstructures of the Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 samples sintered at 1200 °C and 1220 °C, which is 10 °C lower and higher than the above sintering temperature, were also analyzed. Figure 6a shows the SEM micrographs of the Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 powders after additional high-energy ball-milling for 6 h, and Figure 6b,c shows the microstructures of the samples sintered at 1200 °C and 1220 °C. The sintered samples exhibit dense microstructures with no AGG grains. Low-magnification images of these samples are presented in Figure 6c–e. It is a very encouraging that such a dense microstructure was obtained even at 1200 °C sintering, which is 20–30 °C lower than the general sintering temperature for hexaferrite magnets, and this without the use of sintering additives.
Finally, anisotropic sintered magnets with the optimal multi-cation composition, Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05, were fabricated at a commercial ferrite magnet manufacturing site with pilot-level processing machines. Here, the differences from fabricating isotropic magnets on a laboratory scale were that the amount of raw material powder used was scaled up by 100 times, the fine milling process was improved, and a wet magnetic-field molding method was applied during the pressing process. The magnetic hysteresis curves, 4πM-H and B-H, for the anisotropic magnets of Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 sintered at 1200 °C and 1220 °C with and without sintering additives are presented in Figure 7b–e, along with those of a non-substituted SrFe12O19 (SFO) magnet, shown in Figure 7a, for comparison. The SFO sample was a general-grade commercial ferrite magnet manufactured under optimal process conditions. The sintering conditions, sintered sample density, and magnet parameters including 4πMS, iHC, bHC, Br, and BHmax are listed in Table 3, where iHC and bHC are the absolute values of H when 4πM = 0 and B = 0, respectively, in the 4πM-H and B-H curves. Refer to Table 3 for process conditions, such as sintering additives and sintering temperature, for each sample ID. The demagnetization curves of all the samples are plotted in Figure 7f. First, when comparing the overall demagnetization curve of the isotropic magnet without the magnetic-field molding process, as shown in Figure 4d, the squareness and Br values increased significantly. It can also be seen that the performance of the Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 magnets significantly improved compared to the SFO magnets. The best performance of BHmax = 4.70 M·G·Oe was obtained when the Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 sample was sintered at 1200 °C using sintering additives of 0.4 wt.% SiO2 + 0.8 wt.% CaCO3. It is very promising that high magnet performance can be achieved even at a significantly lower sintering temperature, and that Co can be partly replaced with the much less expensive Mn.

4. Conclusions

Three series of samples with the chemical formulas of SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr0.4Ca0.3La0.3Fe11.7-yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) were prepared sequentially, and their hard-magnetic properties were evaluated to determine the optimal composition. The optimal Si-Mg doping level was x = 0.05 in the first series of samples, SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), because it increased both saturation magnetization (MS) and coercivity (HC) and did not produce any second phase at the sintering temperature (TS) of 1230 °C. To further improve the characteristics, cation compositions of Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05 (y = 0, 1, 1.5, 2) were designed by combining La0.4Ca0.3Co0.2 and Si0.05Mg0.05 substitutions while controlling the Fe content. The optimal y value was determined to be 1.5 as it resulted in the highest residual magnetic flux density (Br), and a pure M-type phase without a second phase was confirmed at Fe-deficient y ≥ 1.5. Finally, the effect of the Mn doping level z in Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) samples was evaluated. The doping with Mn (z = 0.05, 0.1, 0.2) caused abnormal grain growth and decreased the coercivity at TS = 1230 °C. When TS was decreased to 1210 °C and the ball-milling process was improved, further enhanced hard-magnetic properties were obtained for the z = 0.1 sample, with relatively high Br and HC values simultaneously. Thus, the multi-cation composition optimized in this study was Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19, and when an anisotropic magnet at the pilot level was produced with this composition, a permanent ferrite magnet with excellent performance of Br = 4.42 kG, HC = 3.57 kOe, and BHmax = 4.70 M·G·Oe values was achieved. These research results are very promising because they show that it is possible to manufacture magnets with excellent performance at greatly reduced sintering temperatures while also reducing the Co content.

Author Contributions

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

Funding

This study was supported by the Research Fund of the Basic Science Research Program through the National Research Foundation of Korea, funded by the Ministry of Science and ICT (Grant No. 2022R1F1A1062933).

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. 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]
  2. Smit, J.; Wijn, H.P.J. Ferrites: Physical Properties of Ferromagnetic Oxides in Relation to Their Technical Applications; Philips Technical Library: Eindhoven, The Netherlands, 1959. [Google Scholar]
  3. Taguchi, H. Recent improvements of ferrite magnets. Phys. IV Fr. 1997, 7, C1-299–C1-302. [Google Scholar] [CrossRef]
  4. 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. B 2000, 68, 182–185. [Google Scholar] [CrossRef]
  5. 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]
  6. Kang, Y.-M. High saturation magnetization in La–Ce–Zn–doped M-type Sr-hexaferrites. Ceram. Int. 2015, 41, 4354–4359. [Google Scholar] [CrossRef]
  7. 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]
  8. Godara, S.K.; sneh; Kaur, V.; Malhi, P.S.; Ahmed, J.; Alshehri, S.M.; Singh, M.; Verma, S.; Singh, C.; Maji, P.K.; et al. Sol-gel auto-combustion synthesis of double metal-doped barium hexaferrite nanoparticles for permanent magnet applications. J. Sol. Stat. Chem. 2022, 312, 123215. [Google Scholar] [CrossRef]
  9. 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]
  10. Yasmin, N.; Mirza, M.; Muhammad, S.; Zahid, M.; Ahmad, M.; Awan, M.S.; Muhammad, A. Influence of samarium substitution on the structural and magnetic properties of M-type hexagonal ferrites. J. Magn. Magn. Mater. 2018, 446, 276–281. [Google Scholar] [CrossRef]
  11. Yang, Y.; Wang, F.; Shao, J.; Huang, D.; He, H.; Trukhanov, A.V.; Trukhanov, S.V. Influence of Nd-NbZn co-substitution on structural, spectral and magnetic properties of M-type calcium-strontium hexaferrites Ca0.4Sr0.6-xNdxFe12.0-x(Nb0.5Zn0.5)xO19. J. Alloy. Compd. 2018, 765, 616–623. [Google Scholar] [CrossRef]
  12. Huang, K.; Yu, J.; Zhang, L.; Xu, J.; Yang, Z.; Liu, C.; Wang, W.; Kan, X. Structural and magnetic properties of Gd–Zn substituted M-type Ba–Sr hexaferrites by sol-gel auto-combustion method. J. Alloy. Compd. 2019, 803, 971–980. [Google Scholar] [CrossRef]
  13. Almessiere, M.A.; Slimani, Y.; Güner, S.; van Leusen, J.; Baykal, A.; Kögerler, P. Effect of Nb3+ ion substitution on the magnetic properties of SrFe12O19 hexaferrites. J. Mater. Sci. Mater. Electron. 2019, 30, 11181–11192. [Google Scholar] [CrossRef]
  14. Todkar, G.B.; Kunale, R.A.; Kamble, R.N.; Batoo, K.M.; Ijaz, M.F.; Imran, A.; Hadi, M.; Raslan, E.H.; Shirsath, S.E.; Kadam, R.H. Ce–Dy substituted barium hexaferrite nanoparticles with large coercivity for permanent magnet and microwave absorber application. J. Phys. D. Appl. Phys. 2021, 54, 294001. [Google Scholar] [CrossRef]
  15. Shirsath, S.E.; Kadam, R.H.; Batoo, K.M.; Wang, D.; Li, S. Co–Al-substituted strontium hexaferrite for rare earth free permanent magnet and microwave absorber application. J. Phys. D Appl. Phys. 2021, 54, 024001. [Google Scholar] [CrossRef]
  16. Han, G.; Sui, R.; Yu, Y.; Wang, L.; Li, M.; Li, J.; Liu, H.; Yang, W. Structure and magnetic properties of the porous Al-substituted barium hexaferrites. J. Magn. Magn. Mater. 2021, 528, 167824. [Google Scholar] [CrossRef]
  17. Manglam, M.K.; Shukla, A.; Mallick, J.; Yadav, M.K.; Kumari, S.; Zope, M.; Kar, M. Enhancement of coercivity of M-type barium hexaferrite by Ho doping. Mater. Today Proc. 2022, 59, 149–152. [Google Scholar] [CrossRef]
  18. 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]
  19. Ogata, Y.; Takami, T.; Kubota, Y. Development of La-Co Substituted Ferrite Magnets. J. Jpn. Soc. Powder Powder Metall. 2003, 50, 636–641. [Google Scholar] [CrossRef] [Green Version]
  20. Kobayashi, Y.; Hosokawa, S.; Oda, E.; Toyota, S. Magnetic properties and composition of Ca-La-Co M-type ferrites. J. Jpn. Soc. Powder Powder Metall. 2008, 55, 541–546. [Google Scholar] [CrossRef] [Green Version]
  21. Yoo, 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. [Google Scholar] [CrossRef]
  22. Moon, K.-S.; Lim, E.-S.; Kang, Y.-M. Effect of Ca and La substitution on the structure and magnetic properties of M-type Sr-hexaferrites. J. Alloy. Compd. 2019, 771, 350–355. [Google Scholar] [CrossRef]
  23. Yu, P.-Y.; Kim, M.-H.; Kang, Y.-M. Development of a high-performance ferrite magnet fabrication process without sintering additives. Korean J. Met. Mater. 2021, 59, 551–559. [Google Scholar] [CrossRef]
  24. 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]
  25. Cullity, B.D.; Graham, C.D. Introduction to Magnetic Materials, 2nd ed.; Hohn Wiley & Sons: Hoboken, NJ, USA, 2009; p. 13. [Google Scholar]
Figure 1. (ac) XRD patterns of SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) samples sintered at 1230 °C.
Figure 1. (ac) XRD patterns of SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2), Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) samples sintered at 1230 °C.
Applsci 13 00295 g001
Figure 2. (a) Magnetic hysteresis curves of the SrFe12−2xSixMgxO19 (x = 0.0, 0.05, 0.1, 0.2) samples sintered at 1230 °C and (b) plots of MS and HC vs. x.
Figure 2. (a) Magnetic hysteresis curves of the SrFe12−2xSixMgxO19 (x = 0.0, 0.05, 0.1, 0.2) samples sintered at 1230 °C and (b) plots of MS and HC vs. x.
Applsci 13 00295 g002
Figure 3. SEM micrographs of (ad) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) and (eh) Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2) samples sintered at 1230 °C.
Figure 3. SEM micrographs of (ad) SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) and (eh) Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2) samples sintered at 1230 °C.
Applsci 13 00295 g003
Figure 5. (al) SEM micrographs of Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (y = 0, 1, 1.5, 2) samples sintered at 1210 °C and 1230 °C.
Figure 5. (al) SEM micrographs of Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (y = 0, 1, 1.5, 2) samples sintered at 1210 °C and 1230 °C.
Applsci 13 00295 g005
Figure 6. (af) SEM micrographs of Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 (y = 1.0) sintered samples: (a,d) after an additional 6 h of ball-milling and after subsequent sintering at (b,e) 1200 °C and (c,f) 1220 °C at low and high magnification.
Figure 6. (af) SEM micrographs of Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 (y = 1.0) sintered samples: (a,d) after an additional 6 h of ball-milling and after subsequent sintering at (b,e) 1200 °C and (c,f) 1220 °C at low and high magnification.
Applsci 13 00295 g006
Figure 7. Plots of magnetization (4πM) and magnetic flux density (B) vs. H for anisotropic magnets with compositions of (a) SrFe12O19 and (be) Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 sintered at 1200 °C and 1220 °C with and without sintering additives. (f) Demagnetization (4πM vs. H) curves of all the magnets.
Figure 7. Plots of magnetization (4πM) and magnetic flux density (B) vs. H for anisotropic magnets with compositions of (a) SrFe12O19 and (be) Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 sintered at 1200 °C and 1220 °C with and without sintering additives. (f) Demagnetization (4πM vs. H) curves of all the magnets.
Applsci 13 00295 g007
Table 1. Sintered sample densities (ρ), saturation magnetization (MS), and coercivity (HC) values of the SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) samples sintered at 1230 °C. The MS value is defined as the magnetization (M) value at H = 25 kOe.
Table 1. Sintered sample densities (ρ), saturation magnetization (MS), and coercivity (HC) values of the SrFe12−2xSixMgxO19 (x = 0, 0.05, 0.1, 0.2) samples sintered at 1230 °C. The MS value is defined as the magnetization (M) value at H = 25 kOe.
Cation Compositionxρ (g/cm3)MS (emu/g)HC (Oe)
SrFe12−2xSixMgx0.02.7572.083562
0.052.8772.153912
0.12.9170.463402
0.24.0766.492986
Table 2. Sintering temperature (TS), sintered sample density (ρ), remanence magnetic flux density (Br), and coercivity (HC) of the hexaferrites Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) samples.
Table 2. Sintering temperature (TS), sintered sample density (ρ), remanence magnetic flux density (Br), and coercivity (HC) of the hexaferrites Sr0.4Ca0.3La0.3Fe11.7−yCo0.2Si0.05Mg0.05O19 (y = 0, 1, 1.5, 2), and Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.05O19 (z = 0, 0.05, 0.1, 0.2) samples.
Cation CompositionTS (°C)y, z, tρ (g/cm3)Br (kG)HC (kOe)
Sr0.4Ca0.3La0.3Fe11.7-yCo0.2Si0.05Mg0.051230y = 0.04.862.373.78
1230y = 1.04.822.533.91
1230y = 1.54.892.693.05
1230y = 2.05.002.641.89
Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.051230z = 0.04.892.693.05
1230z = 0.054.942.251.41
1230z = 1.04.932.401.35
1230z = 2.04.762.422.85
Sr0.4Ca0.3La0.3Fe10.2Co0.2−zMnzSi0.05Mg0.051210z = 04.772.643.17
1210z = 0.054.762.603.89
1210z = 1.04.732.534.15
1210z = 2.04.642.443.50
Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05 (z = 1.0)1210t = 2 h4.812.493.97
1210t = 4 h4.882.534.00
1210t = 6 h4.922.603.95
The error bar for MS is ± 0.2% and for HC is ±1% of the original values.
Table 3. Sintered sample density (ρ), saturation magnetization (4πMS), intrinsic coercivity (kOe), normal coercivity (bHC), remanent magnetization (Br), and maximum energy product (BHmax) of hexaferrites magnets with composition of SrFe12O19 (SFO) and Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 sintered at 1200 °C and 1220 °C with and without sintering additives, respectively.
Table 3. Sintered sample density (ρ), saturation magnetization (4πMS), intrinsic coercivity (kOe), normal coercivity (bHC), remanent magnetization (Br), and maximum energy product (BHmax) of hexaferrites magnets with composition of SrFe12O19 (SFO) and Sr0.4Ca0.3La0.3Fe10.2Co0.1Mn0.1Si0.05Mg0.05O19 sintered at 1200 °C and 1220 °C with and without sintering additives, respectively.
Magnet IDSintering AdditivesTS
(°C)
ρ (g/cm3)4πMS
(kG)
iHC
(kOe)
bHC
(kOe)
Br
(kG)
BHmax
(M·G·Oe)
SFOOptimized12304.924.393.183.004.214.24
No_1220No additives12205.044.553.312.894.414.44
No_120012005.014.513.403.174.374.55
Ad_1220SiO2 0.4wt% +
CaCO3 0.8wt%
12205.004.543.533.214.404.64
Ad_120012005.014.543.573.294.424.70
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lim, J.-P.; Kang, M.-G.; Kang, Y.-M. Development of Multi-Cation-Doped M-Type Hexaferrite Permanent Magnets. Appl. Sci. 2023, 13, 295. https://doi.org/10.3390/app13010295

AMA Style

Lim J-P, Kang M-G, Kang Y-M. Development of Multi-Cation-Doped M-Type Hexaferrite Permanent Magnets. Applied Sciences. 2023; 13(1):295. https://doi.org/10.3390/app13010295

Chicago/Turabian Style

Lim, Jun-Pyo, Min-Gu Kang, and Young-Min Kang. 2023. "Development of Multi-Cation-Doped M-Type Hexaferrite Permanent Magnets" Applied Sciences 13, no. 1: 295. https://doi.org/10.3390/app13010295

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