Next Article in Journal
Optimizing Gold Nanoparticle Size and Shape for the Fabrication of SERS Substrates by Means of the Langmuir–Blodgett Technique
Next Article in Special Issue
Solvothermal Synthesis Combined with Design of Experiments—Optimization Approach for Magnetite Nanocrystal Clusters
Previous Article in Journal
Nonlinear Optical Characterization of 2D Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Mg and Sb Substitution on the Magnetic Properties of Magnetic Field Annealed MnBi Alloys

1
Powder & Ceramics Division, Korea Institute of Materials Science, Changwon 51508, Korea
2
School of Materials Science and Engineering, Pusan National University, Busan 46241, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(11), 2265; https://doi.org/10.3390/nano10112265
Submission received: 20 October 2020 / Revised: 4 November 2020 / Accepted: 13 November 2020 / Published: 16 November 2020
(This article belongs to the Special Issue Synthesis and Properties of Nanocrystalline Magnetic Materials)

Abstract

:
Rare-earth-free permanent magnets have attracted considerable attention due to their favorable properties and applicability for cost-effective, high-efficiency, and sustainable energy devices. However, the magnetic field annealing process, which enhances the performance of permanent magnets, needs to be optimized for different magnetic fields and phases. Therefore, we investigated the effect of composition on the crystallization of amorphous MnBi to the ferromagnetic low-temperature phase (LTP). The optimal compositions and conditions were applied to magnetic field annealing under 2.5 T for elemental Mg- and Sb/Mg pair-substituted MnBi. The optimum MnBi composition for the highest purity LTP was determined to be Mn56Bi44, and its maximum energy product, (BH)max, was 5.62 MGOe. The Mg-substituted MnBi exhibited enhanced squareness (Mr/Ms), coercivity (Hc), and (BH)max values up to 0.8, 9659 Oe, and 5.64 MGOe, respectively, whereas the same values for the Sb/Mg pair-substituted MnBi were 0.76, 7038 Oe, and 5.60 MGOe, respectively. The substitution effects were also investigated using first-principles calculations. The density of states and total magnetic moments of Mn16Bi15Mg and Mn16Bi15Sb were similar to those of pure Mn16Bi16. Conversely, the Sb-substituted MnBi resulted in a dramatic enhancement in the anisotropy constant (K) from a small negative value (−0.85 MJ/m3) to a large positive value (6.042 MJ/m3).

1. Introduction

Rare-earth-free permanent magnets featuring high maximum energy products ((BH)max) between those of rare-earth and traditional rare-earth-free permanent magnets are in high demand for cost-effective, high-efficiency, and sustainable energy devices [1,2]. Among rare-earth-free permanent magnetic materials, MnBi exhibits the most promising magnetic properties, featuring high (BH)max values that exceed 9 MGOe for powder [3,4,5,6,7] and 7 MGOe for bulk [7,8,9,10]. Notably, the theoretical (BH)max of MnBi is 17.7 MGOe at 300 K [11]. During compaction of the powder samples, the reduced coercivity (Hc) degrades the (BH)max for bulk MnBi. Consequently, the addition of secondary elements or materials, such as Ga (e.g., Mn55Bi44Ga) [12], Sn (e.g., MnBi1-xSnx) [13], NaCl or C [14], and nanoparticles [15], to MnBi has been investigated in efforts to increase the Hc. Furthermore, magnetic field annealing has been used to enhance the magnetic properties for permanent magnet applications [16,17,18,19,20,21].
Magnetic field annealing, a process widely used to fabricate Alnico magnets [22], induces aligned texturing in the microstructure. However, the process is not efficient for all permanent magnets owing to variations in the field dependence of the magnetic phase crystallization. Secondary elements such as Sn, Sb, Mg, and In have been substituted/doped to enhance the effectiveness of field annealing for MnBi permanent magnetic materials [23,24,25,26,27]. The magnetic properties of the secondary element-doped/substituted MnBi are superior to those of MnBi fabricated using a field annealing process, revealing higher magnetization, Hc, and squareness (Mr/Ms) than those of binary MnBi [23,24,25,26,27]. Therefore, the highest (BH)max (12 MGOe) was achieved for Mn50Bi46Mg3In0.5Sb0.5 bulk magnetic material [27].
In this study, we systematically investigated the composition (Mn52+xBi48−x, 0 ≤ x ≤ 8) and annealing condition dependencies of the purity of the low-temperature phase (LTP). The optimized composition was used in magnetic field annealing experiments under the optimized annealing conditions with an applied field of 2.5 T. The dependence of the magnetic properties on the substitution of Mg and Sb/Mg pairs over large composition ranges, i.e., Mn56Bi44−xMgx (0 ≤ x ≤ 5) and Mn56Bi43.5−xSb0.5Mgx (0 ≤ x ≤ 5), were also systematically investigated. To better comprehend the experimental data, the substitution effects were analyzed using total energies based on different spin configurations, densities of states (DOSs), and magnetic moments.

2. Experimental and Calculations

MnBi, Mn-Bi-Mg, and Mn-Bi-Sb-Mg bulks were prepared using the following procedure: melt spinning, cold pressing, and magnetic annealing. The commercial raw materials of Mn (99.95%), Bi (99.999%), Mg (99.98%), and Sb (99.999%) were mixed in the desired atomic ratios of Mn52+xBi48−x (0 ≤ x ≤ 8), Mn56Bi44−xMgx (0 ≤ x ≤ 5), and Mn56Bi43.5−xSb0.5Mgx (0 ≤ x ≤ 5). The mixtures were melted in an Ar environment in quartz nozzles and ejected through a 0.4 mm orifice using pressurized Ar gas onto a copper wheel rotating at a velocity of 40 m/s. The melt-spun ribbons were manually ground (5 min) in ethyl alcohol and then pressed into pellets in tungsten carbide molds (500 MPa). The produced pellets were cut into small pieces (1.5 mm × 1.0 mm × 3 mm) and subjected to magnetic annealing under a magnetic field of 2.5 T. The annealing temperature was optimized by measuring the temperature dependence of the magnetic moment of an amorphous bulk, where the ramping rate was maintained at 16 °C/min. Once the temperature reached the target temperature, the heating element was immediately removed to rapidly cool the samples.
The crystalline structures and phase purities of the products were characterized using X-ray diffraction (XRD) with Cu Kα (λ = 1.5406 Å) radiation (Rigaku Inc., D/Max-2500VL/PC, Tokyo, Japan). Further, the measured XRD patterns were analyzed using the Rietveld refinement method with the program FullProf. Magnetic annealing and magnetic property measurements were performed using a vibrating sample magnetometer (VSM) (Microsense, EZ9, Lowell, MA, USA). The magnetic properties of the bulk samples were corrected according to their shape with a demagnetization factor of 0.1655.
The WIEN2k package [28] was used to perform first-principles calculations. The package is based on density functional theory (DFT) and uses the full-potential linearized augmented plane wave method. The 3p, 3d, and 4s states of Mn; 4d, 4f, 5p, 5d, 6s, and 6p states of Bi; 3s state of Mg; and 4p, 4d, 5s, and 5p states of Sb were treated as valence states in the calculations. Unit cell and supercell calculations used 28 × 28 × 24 and 25 × 25 × 15 reciprocal space meshes, which generate 18,816 and 9375 k-points in the irreducible part of the Brillouin zone, respectively. The total energies were calculated within an accuracy of 0.0013 eV. The experimental lattice constants (a = 4.256 Å and c = 6.042 Å at 4.2 K) [29] in Figure 1 were used, and the muffin-tin radii (RMT) for Mn, Bi, Mg, and Sb were set as 2.5 a.u. The cutoff parameter of PMT × Kmax = 7.0 with lmax = 10 inside the muffin-tin spheres was used to expand the wavefunctions in terms of the lattice harmonics. All spin-polarized and spin-orbit coupling calculations were based on DFT within the local-spin-density approximation (LSDA).

3. Results and Discussion

The degree of amorphization plays an important role in crystallizing LTP Mn50Bi50 in the direction of the magnetic field. Therefore, we prepared Mn52+xBi48x (0 ≤ x ≤ 8) samples to investigate the composition dependence of the degree of amorphization and their crystallization to LTP Mn50Bi50 after annealing at 300 °C for 24 h. Melt-spun Mn52+xBi48x (0 ≤ x ≤ 8) products contain small amounts of droplet-shaped particles and ribbons. These particles were removed, and only the ribbons were collected to fabricate bulk samples. It is noted that Mn and Bi were partially reacted with quartz nozzle during the melt-spinning process; therefore, the final compositions converge into Mn50Bi50 regardless of the initial compositions. The bulks were loaded into a vacuum tube furnace and annealed. Figure 2a shows the XRD patterns of the Mn52+xBi48x (0 ≤ x ≤ 8) melt-spun ribbons. All the samples exhibited minor peaks of Bi, whereas LTP Mn50Bi50 was not detected. Therefore, the melt-spun ribbons mostly comprised an amorphous phase with a small quantity of Bi crystallites. The differences in the peak intensities of Bi for the five samples were indistinguishable, except for Mn54Bi46. Figure 2b shows the XRD patterns of the annealed Mn52+xBi48x (0 ≤ x ≤ 8) bulk samples. The purities of the samples were initially enhanced with increasing x, but they were not enhanced significantly when x ≥ 4; the calculated purities were 84%, 93%, 96%, 96%, and 97% for x = 0, 2, 4, 6, and 8, respectively. Therefore, we used the composition x = 4, i.e., Mn56Bi44, for the following experiment.
The annealing temperature of amorphous Mn56Bi44 was optimized using VSM. Figure 3 depicts the temperature dependence of the magnetic moment per Mn atom of an amorphous Mn56Bi44 bulk sample under an applied magnetic field of 2.5 T. The amorphous Mn56Bi44 bulk sample was heated from 130 to 550 °C while its magnetic moment was measured using VSM. The measured magnetic moment of the paramagnetic amorphous bulk sample initially decreased with increasing temperature but began increasing at 220 °C due to crystallization of the amorphous phase. The highest magnetic moment was measured at temperatures ranging from 310 to 340 °C. It is noted that further heating of the sample above 355 °C results in irreversible decomposition of LTP Mn50Bi50 into Mn52Bi48 + Bi [30]. Therefore, we selected an annealing temperature of 320 °C for the annealing process. Notably, the magnetic moment declined at temperatures between 430 and 460 °C, with a sharp decrease at 450 °C. This temperature is considerably higher than the phase transformation temperature (340–360 °C) of MnBi from LTP to the high-temperature phase (HTP) [29]. This temperature disagreement at high temperatures may be attributed to heat loss in the open heating circuit of the VSM system.
Figure 4 depicts the refined XRD pattern of the Mn56Bi44 bulk sample annealed at 320 °C for 5 min. The phase identity and weight fraction were analyzed using the Rietveld refinement method with the program FullProf. The refined XRD pattern confirmed that the annealed Mn56Bi44 bulk had a hexagonal structure with a space group of P63/mmc, and the lattice constants a0 and c0 were 4.285 ± 0.001 and 6.113 ± 0.001 Å, respectively. The annealed Mn56Bi44 bulk mainly comprised the LTP, and the peak at approximately 27.24° corresponds to an impure Bi phase. After heat treatment at 320 °C for 5 min, the purity of the LTP Mn50Bi50 increased to 98.15 wt.%. The Mn and Bi atoms occupied the 2a site (0, 0, 0) and the 2c site (1/3, 2/3, 1/4), respectively [31].
The magnetic hysteresis loops of the Mn56Bi44xMgx (0 ≤ x ≤ 5) bulk samples obtained after magnetic annealing under 2.5 T are shown in Figure 5. The hysteresis loops were smooth and did not contain any kinks until x = 3; however, a kink appeared at x = 4. As the Mg content increased, the magnetization at the maximum magnetic field and remanent magnetization decreased, whereas Hc increased to 9659 Oe for x = 4 and then decreased at higher values of x. The magnetization, squareness, Hc, and (BH)max of the Mn56Bi44xMgx (0 ≤ x ≤ 5) bulk samples are plotted in Figure 6. Although the magnetization at the maximum field and remanent magnetization decreased as x increased, (BH)max increased slightly from 5.62 MGOe (for Mn56Bi44) to 5.64 MGOe (for Mn56Bi43Mg1 and Mn56Bi42Mg2) because of the increased squareness (Mr/Ms ≈ 0.8). At x values exceeding 2, (BH)max decreased linearly. The decreasing magnetization and increasing Hc for x ≥ 2 originate from phase decomposition into nonmagnetic phases. The XRD patterns for the field-annealed samples will be published elsewhere. These impurities act as pinning centers for domain walls by strongly inhibiting their motion, thus reducing the efficiency of this reversal mechanism.
Sb-substituted Mn-Bi-Mg, i.e., Mn56Bi43.5xSb0.5Mgx (0 ≤ x ≤ 5) bulk, was also fabricated to investigate the effect of Sb/Mg pair substitution on the magnetic properties. The magnetic hysteresis loops of the Mn56Bi43.5xSb0.5Mgx (0 ≤ x ≤ 5) bulk samples are shown in Figure 7. The magnetic annealing process and conditions were the same as those for Mn56Bi44xMgx (0 ≤ x ≤ 5). The magnetization at maximum field and remanent magnetization decreased with increasing Mg content, while Hc linearly increased until x = 4, reaching 7038 Oe, as shown in Figure 8. The hysteresis loops were smooth, containing no kinks until x = 3, but a kink appeared at x = 4. Substitution of a small amount of Sb into Mn-Bi-Mg increased Hc until x = 3 but caused severe phase decomposition at higher Mg contents; the LTP was completely decomposed at x = 5. The increased (BH)max for Mn56Bi44xMgx (0 ≤ x ≤ 5) is attributed to the enhanced squareness. However, the squareness monotonically decreased with increasing Mg content for Mn56Bi43.5xSb0.5Mgx (0 ≤ x ≤ 5), resulting in decreased (BH)max. The decreasing magnetization and increasing Hc for x ≥ 3 are attributed to the existence of nonmagnetic secondary phases, which act as pinning centers for the domain walls.
The substitution effects of Mg and Sb on the magnetic properties of MnBi were also investigated using first-principles calculations. Mg and Sb are presumed to substitute in the Bi site in the unit cell of MnBi (Figure 1). The crystal structures of Mn16Bi15Mg (or Sb) and Mn16Bi14SbMg for the calculations are shown in Figure 9. Figure 10 shows the DOSs of LTP MnBi, Mn16Bi15Mg, Mn16Bi15Sb, and Mn16Bi14SbMg, which were calculated based on spin-polarized and spin-orbit coupling. The s bands of Bi, Mg, and Sb mainly contributed to the low-energy region below the Fermi energy (EF), whereas the d band of Mn mainly contributed to the high-energy region above EF, as seen in Figure 10. The DOS is a highly degenerate energy state that is slightly below and above EF in the majority and minority spin states, respectively. These highly degenerate energy states near the EF are the origin of the magnetic moment of LTP MnBi-based magnetic materials because they contribute to the difference between the number of electrons in the majority and minority spin states below the EF. In contrast to the highly degenerate DOSs below (above) the EF for the majority (minority) spin states of LTP MnBi, Mn16Bi15Mg, Mn16Bi15Sb, and Mn16Bi14SbMg (see Figure 10), the EF lies at the low DOS between the highly degenerated spin states for the majority and minority spin states. This EF position is the origin of the phase stability of the above-mentioned LTP MnBi-based magnetic materials.
The spin magnetic moments for Mn, Bi, Mg, and Sb and the anisotropy constants (K) for MnBi, MnMg, Mn16Bi15Mg, Mn16Bi15Sb, and Mn16Bi14SbMg are given in Table 1. The spins of the Mn atoms in MnMg were antiparallel to each other, resulting in a small total magnetic moment. The magnetic moment of Mn was approximately 3.57 μB, whereas those of Bi, Mg, and Sb were approximately −0.1 μB for all ferromagnetic MnBi-based materials, as shown in Table 1. Therefore, the magnetic properties of MnBi-based magnetic materials are mostly dependent on the Mn atoms. The total magnetic moments of a Mg or Sb element or a Mg/Sb pair-substituted MnBi were not degraded; the total magnetic moments of Mn16Bi15Mg, Mn16Bi15Sb, and Mn16Bi14SbMg were similar to that of MnBi (56.352 μB = 7.044 μB × 8 unit cells).
The K value of MnBi was calculated to be −0.850 × 106 J/m3, which corresponds to an in-plane spin direction. The spins of MnBi at 0 K lie in the basal plane but rotate to the c-axis at approximately 90 K [9]. The K of Mg-substituted MnBi, i.e., Mn16Bi15Mg, was almost identical to that of pure MnBi, whereas that of Sb-substituted MnBi, i.e., Mn16Bi15Sb, was much larger than Mn16Bi15Mg in the c-axis, as shown in Table 1. Therefore, one Mg element substitution into Mn16Bi16 resulted in no K change, whereas one Sb element substitution resulted in an out-of-plane spin direction; this property is necessary for the development of permanent magnets. Moreover, the K of 6.042 × 106 J/m3 for Mn16Bi15Sb is comparable to that of commercial rare-earth permanent magnets, such as 4.9 × 106 J/m3 for Nd2Fe14B and 4.2 × 106 J/m3 for Sm2Co17 [32]. In addition, the K of the Sb/Mg pair-substituted MnBi, i.e., Mn16Bi14SbMg, was larger than that of Mn16Bi15Mg due to the substitution effect of Sb.

4. Conclusions

We optimized the initial Mn-Bi composition and annealing temperature for high-purity LTP using melt-spinning and annealing processes. The optimized annealing conditions were applied to fabricate Mg and Mg- and Sb-substituted Mn-Bi using magnetic annealing. The magnetic field annealing of Mn56Bi44 resulted in an enhanced squareness of 0.74, which was further increased to 0.8 upon Mg substitution. The highest coercivity and maximum energy product of Mg-substituted Mn-Bi were 9659 Oe (for Mn56Bi40Mg4) and 5.64 MGOe (for Mn56Bi42Mg2), respectively. First-principles calculations indicated that the total magnetic moments of Mn16Bi15Mg and Mn16Bi15Sb were similar to the total magnetic moment of pure Mn16Bi16. In contrast, the anisotropy constant K changed from −0.85 to 6.042 MJ/m3 when Sb was substituted, indicating that the spin direction in the basal plane changed to the c-axis.

Author Contributions

Conceptualization, J.T.L.; Formal analysis, J.-W.K.; Investigation, H.-D.Q. and Y.Y.; Supervision, C.-J.C.; Writing—original draft, J.P.; Writing—review & editing, J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly supported by the Future Materials Discovery Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (2016M3D1A1027835) and by the Fundamental Research Program (PNK6940) of the Korea Institute of Materials Science (KIMS).

Conflicts of Interest

The authors declare no conflict of interest. The sponsors had no involvement in the study design; in the collection, analysis, or interpretation of data; in the writing of the report; or in the decision to submit the article for publication.

References

  1. Vagati, A. The Synchronous Reluctance Solution: A New Alternative in AC drives. In Proceedings of the IECON’94–20th Annual Conference of IEEE Industrial Electronics, Bologna, Italy, 5–9 September 1994; Volume 1, pp. 1–13. [Google Scholar]
  2. Lipo, T.A. Synchronous reluctance machines-a viable alternative for AC drives? Electr. Mach. Power Syst. 1991, 19, 659–671. [Google Scholar] [CrossRef]
  3. Rama Rao, N.V.; Gabay, A.M.; Hadjipanayis, G.C. Anisotropic fully dense MnBi permanent magnet with high energy product and high coercivity at elevated temperatures. J. Phys. D Appl. Phys. 2013, 46, 062001. [Google Scholar] [CrossRef]
  4. Cao, J.; Huang, Y.L.; Hou, Y.H.; Shi, Z.Q.; Yan, X.T.; Zhong, Z.C.; Wang, G.P. Microstructure and magnetic properties of MnBi alloys with high coercivity and significant anisotropy prepared by surfactant assisted ball milling. J. Magn. Magn. Mater. 2019, 473, 505–510. [Google Scholar] [CrossRef]
  5. Chinnasamy, C.; Jasinski, M.M.; Ulmer, A.; Li, W.; Hadjipanayis, G.; Liu, J. Mn-Bi magnetic powders with high coercivity and magnetization at room temperature. IEEE Trans. Magn. 2012, 48, 3641–3643. [Google Scholar] [CrossRef]
  6. Yang, Y.; Park, J.; Lim, J.T.; Kim, J.-W.; Li, O.L.; Choi, C.-J. Effect of phase purity on enhancing the magnetic properties of Mn-Bi alloy. J. Magn. Magn. Mater. 2021, 517, 167344. [Google Scholar] [CrossRef]
  7. Cui, J.; Choi, J.P.; Li, G.; Polikarpov, E.; Darsell, J.; Overman, N.; Olszta, M.; Schreiber, D.; Bowden, M.; Droubay, T.; et al. Thermal stability of MnBi magnetic materials. J. Phys. Condens. Matter 2014, 26, 064212. [Google Scholar] [CrossRef]
  8. Kim, S.; Moon, H.; Jung, H.; Kim, S.-M.; Lee, H.-S.; Choi-Yim, H.; Lee, W. Magnetic properties of large-scaled MnBi bulk magnets. J. Alloys Compd. 2017, 708, 1245–1249. [Google Scholar] [CrossRef]
  9. Yang, J.B.; Yelon, W.B.; James, W.J.; Cai, Q.; Kornecki, M.; Roy, S.; Ali, N.; L’Heritier, P. Crystal structure, magnetic properties and electronic structure of the MnBi intermetallic compound. J. Phys. Condens. Matter 2002, 14, 6509–6519. [Google Scholar] [CrossRef]
  10. Poudyal, N.; Liu, X.; Wang, W.; Nguyen, V.V.; Ma, Y.; Gandha, K.; Elkins, K.; Liu, J.P.; Sun, K.; Kramer, M.J.; et al. Processing of MnBi bulk magnets with enhanced energy product. AIP Adv. 2016, 6, 056004. [Google Scholar] [CrossRef] [Green Version]
  11. Park, J.; Hong, Y.-K.; Lee, J.; Lee, W.; Kim, S.-G.; Choi, C.-J. Electronic structure and maximum energy product of MnBi. Metals 2014, 4, 455–464. [Google Scholar] [CrossRef] [Green Version]
  12. Yang, Y.; Kim, J.-W.; Si, P.-Z.; Qian, H.-D.; Shin, Y.; Wang, X.; Park, J.; Li, O.L.; Wu, Q.; Ge, H.; et al. Effects of Ga-doping on the microstructure and magnetic properties of MnBi alloys. J. Alloys Compd. 2018, 769, 813–816. [Google Scholar] [CrossRef]
  13. Sakuma, A.; Manabe, Y.; Kota, Y. First principles calculation of magnetocrystalline anisotropy energy of MnBi and MnBi1-xSnx. J. Phys. Soc. Jpn. 2013, 82, 073704. [Google Scholar] [CrossRef]
  14. Qian, H.-D.; Park, J.; Lim, J.T.; Yang, Y.; Si, P.-Z.; Kim, J.W.; Choi, C.-J.; Cho, K.M. Magnetic properties of MnBi bulk magnets with NaCl and C addition. AIP Adv. 2019, 9, 115213. [Google Scholar] [CrossRef]
  15. Wang, X.; Qian, H.; Si, P.-Z.; Yang, Y.; Choi, C.-J.; Park, J.; Wang, X.; Ge, H.; Shinde, K.P.; Chung, K. Structure and magnetic properties of MnBi nanoparticles prepared by laser ablation and arc-discharge method. IEEE Trans. Magn. 2018, 54, 2301005. [Google Scholar] [CrossRef]
  16. Yasuda, H.; Ohnaka, I.; Yamamoto, Y.; Tokieda, K.; Kishio, K. Formation of crystallographically aligned BiMn grains by semi-solid processing of rapidly solidified Bi-Mn alloys under a magnetic field. Mater. Trans. 2003, 44, 2207–2212. [Google Scholar] [CrossRef]
  17. Ren, Z.-M.; Li, X.; Deng, K.; Wang, H.; Zhuang, Y. Solidification structures of Bi-Mn alloys under a high magnetic field. J. Shanghai Univ. 2006, 10, 74–77. [Google Scholar] [CrossRef]
  18. Kobayashi, R.; Mitsui, Y.; Takahashi, K.; Uda, S.; Koyama, K. In-field annealing and quenching for ferromagnetic MnBi under 19 T. IEEE Trans. Magn. 2019, 55, 1000204. [Google Scholar] [CrossRef]
  19. Mitsui, Y.; Abematsu, K.; Umetsu, R.Y.; Takahashi, K.; Koyama, K. Magnetic field effects on liquid-phase reactive sintering of MnBi. J. Magn. Magn. Mater. 2016, 400, 304–306. [Google Scholar] [CrossRef]
  20. Zhang, W.Y.; Kharel, P.; George, T.; Li, X.Z.; Mukherjee, P.; Valloppilly, S.; Sellmyer, D.J. Grain alignment due to magnetic-field annealing in MnBi:Bi nanocomposites. J. Phys. D Appl. Phys. 2016, 49, 455002. [Google Scholar] [CrossRef] [Green Version]
  21. Si, P.Z.; Yang, Y.; Yao, L.L.; Qian, H.D.; Ge, H.L.; Park, J.; Chung, K.C.; Choi, C.J. Magnetic-field-enhanced reactive synthesis of MnBi from Mn nanoparticles. J. Magn. Magn. Mater. 2019, 476, 243–247. [Google Scholar] [CrossRef]
  22. Mishima, T. Magnet Steel Containing Nickel and Aluminium. U.S. Patent 2,027,994, 14 January 1936. [Google Scholar]
  23. Zhang, W.; Balasubramanian, B.; Kharel, P.; Pahari, R.; Valloppilly, S.R.; Li, X.; Yue, L.; Skomski, R.; Sellmyer, D.J. High energy product of MnBi by field annealing and Sn alloying. APL Mater. 2019, 7, 121111. [Google Scholar] [CrossRef] [Green Version]
  24. Gabay, A.M.; Hadjipanayis, G.C.; Cui, J. Effect of Sb substitution on crystal structure, texture and hard magnetic properties of melt-spun MnBi alloys. J. Alloys Compd. 2019, 792, 77–86. [Google Scholar] [CrossRef] [Green Version]
  25. Gabay, A.M.; Hadjipanayis, G.C.; Cui, J. New anisotropic MnBi permanent magnets by field-annealing of compacted melt-spun alloys modified with Mg and Sb. J. Magn. Magn. Mater. 2020, 495, 165860. [Google Scholar] [CrossRef]
  26. Gabay, A.M.; Hadjipanayis, G.C. Effect of Mg content in melt-spun Mn–Bi–Mg–Sb–In alloys on the structure and properties of field-annealed magnets. IEEE Magn. Lett. 2020, 11, 7503304. [Google Scholar] [CrossRef]
  27. Gabay, A.M.; Hadjipanayis, G.C.; Cui, J. Development of rare-earth-free bulk magnets with energy product up to 12 MGOe in field annealed Mn–Bi–Mg–In–Sb alloys. J. Alloys Compd. 2020, 822, 153663. [Google Scholar] [CrossRef]
  28. Blaha, P.; Schwarz, K.; Madsen, G.; Kvasnicka, D.; Luitz, J. WIEN2k: An Augmented Plane Wave Plus Local Orbitals Program for Calculating Crystal Properties; Vienna University of Technology: Vienna, Austria, 2001. [Google Scholar]
  29. Roberts, B.W. Neutron diffraction study of the structures and magnetic properties of manganese bismuthide. Phys. Rev. 1956, 104, 607–616. [Google Scholar] [CrossRef]
  30. Chen, T. Contribution to the equilibrium phase diagram of the Mn-Bi system near MnBi. J. Appl. Phys. 1974, 45, 2358–2360. [Google Scholar] [CrossRef]
  31. Christopher, N.R.; Singh, N.; Singh, S.K.; Gahtori, B.; Mishra, S.K.; Dhar, A.; Awana, V.P.S. Appreciable magnetic moment and energy density in single-step normal route synthesized MnBi. J. Supercond. Novel Magn. 2013, 26, 3161–3165. [Google Scholar] [CrossRef] [Green Version]
  32. Coey, J.M.D. Magnetism and Magnetic Materials, 1st ed.; Cambridge University Press: Cambridge, UK, 2009; pp. 401–404. [Google Scholar]
Figure 1. Hexagonal low-temperature phase (LTP) MnBi. The Mn and Bi atoms are blue and yellow, respectively.
Figure 1. Hexagonal low-temperature phase (LTP) MnBi. The Mn and Bi atoms are blue and yellow, respectively.
Nanomaterials 10 02265 g001
Figure 2. XRD patterns of the (a) ground Mn52+xBi48x melt-spun ribbons and (b) annealed bulk samples (300 °C for 24 h).
Figure 2. XRD patterns of the (a) ground Mn52+xBi48x melt-spun ribbons and (b) annealed bulk samples (300 °C for 24 h).
Nanomaterials 10 02265 g002
Figure 3. Temperature dependence of the magnetic moment per Mn atom of an amorphous Mn56Bi44 bulk sample under an applied magnetic field of 2.5 T.
Figure 3. Temperature dependence of the magnetic moment per Mn atom of an amorphous Mn56Bi44 bulk sample under an applied magnetic field of 2.5 T.
Nanomaterials 10 02265 g003
Figure 4. Refined XRD pattern of the annealed Mn56Bi44 bulk samples (320 °C for 5 min).
Figure 4. Refined XRD pattern of the annealed Mn56Bi44 bulk samples (320 °C for 5 min).
Nanomaterials 10 02265 g004
Figure 5. Magnetic hysteresis loops of Mn56Bi44xMgx (0 ≤ x ≤ 5) bulk samples.
Figure 5. Magnetic hysteresis loops of Mn56Bi44xMgx (0 ≤ x ≤ 5) bulk samples.
Nanomaterials 10 02265 g005
Figure 6. Composition dependence of magnetization, squareness, Hc, and (BH)max for the Mn56Bi44-xMgx (0 ≤ x ≤ 5) bulk samples.
Figure 6. Composition dependence of magnetization, squareness, Hc, and (BH)max for the Mn56Bi44-xMgx (0 ≤ x ≤ 5) bulk samples.
Nanomaterials 10 02265 g006
Figure 7. Magnetic hysteresis loops of Mn56Bi43.5xSb0.5Mgx (0 ≤ x ≤ 5) bulk samples.
Figure 7. Magnetic hysteresis loops of Mn56Bi43.5xSb0.5Mgx (0 ≤ x ≤ 5) bulk samples.
Nanomaterials 10 02265 g007
Figure 8. Composition dependence of magnetization, squareness, Hc, and (BH)max for Mn56Bi43.5xSb0.5Mgx (0 ≤ x ≤ 5) bulk samples.
Figure 8. Composition dependence of magnetization, squareness, Hc, and (BH)max for Mn56Bi43.5xSb0.5Mgx (0 ≤ x ≤ 5) bulk samples.
Nanomaterials 10 02265 g008
Figure 9. Crystal structures of (a) Mn16Bi15Mg (or Sb) and (b) Mn16Bi14SbMg used for the first-principles calculations. The Mn, Bi, Mg, and Sb atoms are in blue, yellow, green, and red, respectively.
Figure 9. Crystal structures of (a) Mn16Bi15Mg (or Sb) and (b) Mn16Bi14SbMg used for the first-principles calculations. The Mn, Bi, Mg, and Sb atoms are in blue, yellow, green, and red, respectively.
Nanomaterials 10 02265 g009
Figure 10. Density of states (DOS) for (a) MnBi, (b) Mn16Bi15Mg, (c) Mn16Bi15Sb, and (d) Mn16Bi14SbMg.
Figure 10. Density of states (DOS) for (a) MnBi, (b) Mn16Bi15Mg, (c) Mn16Bi15Sb, and (d) Mn16Bi14SbMg.
Nanomaterials 10 02265 g010
Table 1. Calculated Mn, Bi, Mg, Sb, interstitial, and total magnetic moments and anisotropy constants.
Table 1. Calculated Mn, Bi, Mg, Sb, interstitial, and total magnetic moments and anisotropy constants.
MaterialsMnBiMgSbInterstitial, Int.Total Magnetic Moments, μBAnisotropy Constant, K (×106 J/m3)
MnBi3.571−0.103--0.1047.044−0.850
MnMg2.331-0.002-0.0060.06719.690
Mn16Bi15Mg3.577−0.108−0.071-0.79056.371−0.898
Mn16Bi15Sb3.573−0.121-−0.1240.74156.2106.042
Mn16Bi14SbMg3.579−0.134−0.071−0.1310.78456.368−0.441
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qian, H.-D.; Yang, Y.; Lim, J.T.; Kim, J.-W.; Choi, C.-J.; Park, J. Effects of Mg and Sb Substitution on the Magnetic Properties of Magnetic Field Annealed MnBi Alloys. Nanomaterials 2020, 10, 2265. https://doi.org/10.3390/nano10112265

AMA Style

Qian H-D, Yang Y, Lim JT, Kim J-W, Choi C-J, Park J. Effects of Mg and Sb Substitution on the Magnetic Properties of Magnetic Field Annealed MnBi Alloys. Nanomaterials. 2020; 10(11):2265. https://doi.org/10.3390/nano10112265

Chicago/Turabian Style

Qian, Hui-Dong, Yang Yang, Jung Tae Lim, Jong-Woo Kim, Chul-Jin Choi, and Jihoon Park. 2020. "Effects of Mg and Sb Substitution on the Magnetic Properties of Magnetic Field Annealed MnBi Alloys" Nanomaterials 10, no. 11: 2265. https://doi.org/10.3390/nano10112265

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