Next Article in Journal
Time-Varying Pilot’s Intention Identification Based on ESAX-CSA-ELM Classification Method in Complex Environment
Next Article in Special Issue
High-Temperature Thermodynamics Modeling of Graphite
Previous Article in Journal
Current Approaches and Applications in Natural Language Processing
Previous Article in Special Issue
Analytic Binary Alloy Volume–Concentration Relations and the Deviation from Zen’s Law
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermodynamics and Magnetism of SmFe12 Compound Doped with Co and Ni: An Ab Initio Study

Critical Materials Institute, Lawrence Livermore National Laboratory, Livermore, CA 94551-0808, USA
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(10), 4860; https://doi.org/10.3390/app12104860
Submission received: 12 April 2022 / Revised: 7 May 2022 / Accepted: 8 May 2022 / Published: 11 May 2022
(This article belongs to the Special Issue Feature Paper Collection in Section Materials)

Abstract

:
Ni-doped Sm(Fe1−xCox)12 alloys are investigated for their magnetic properties. The Sm(Fe,Co)11M1 compound (M acts as a stabilizer) with the smallest (7.7 at.%) rare-earth-metal content has been recognized as a possible contender for highly efficient permanent magnets thanks to its significant anisotropy field and Curie temperature. The early transition metals (Ti-Mn) as well as Al, Si, and Ga stabilize the SmFe12 compound but significantly decrease its saturation magnetization. To keep the saturation magnetization in the range of 1.4–1.6 T, we suggest replacing a certain amount of Fe and Co in the Sm(Fe1−xCox)12 alloys with Ni. Ni plays the role of a thermodynamic stabilizer, and contrary to the above-listed elements, has the spin moment aligned parallel to the spin moment of the SmFe12 compound, thereby boosting its saturation magnetization without affecting the anisotropy field or Curie temperature.

1. Introduction

Three main material parameters determine the intrinsic properties of hard magnetic materials: (i) the saturation magnetization, Ms; (ii) the Curie temperature, Tc; (iii) the anisotropy field, Ha [1]. These three parameters all need to be of a significant value for an economically compelling permanent magnet, i.e., μ0Ms ~≥ 1.25 T, Tc ~≥ 550 K, and μ0Ha ~≥ 3.75 T (μ0 is the vacuum permeability). By mixing transition-metal (TM) with rare-earth-metal (RE) atoms in various intermetallic compounds [1,2], one can produce a material with these important magnetic properties.
The advancement of Nd-Fe-B sintered magnets in the 1980′s by Sagawa et al. [3,4,5,6] led to the introduction of iron-containing compounds that were of interest due to their considerable magnetic moments and anisotropies, which are necessary for industrially functional permanent magnetic materials [7,8,9,10,11,12]. Currently, the Nd2Fe14B1-based sintered magnets (or so-called Neomax magnets) are considered the technological standard due to their robust performance as a result of their considerable saturation magnetization (µ0MS = 1.61 T, anisotropy field, µ0Ha = 7.6 T, sufficient coercivity, µ0Hc ≥ 0.8 T) at room temperature [1,3,4,5,6,7,8,9,10,11,12,13] and enormous maximum energy production (|BH|max~515 kJ/m3) [14]. A deficiency of the Nd2Fe14B1 magnets is their low Curie temperature, Tc = 588 K, and the notable deterioration of the anisotropy field with temperature [12]. Additionally, as was mentioned by Goll et al. [15], the corrosion resistance of the Nd2Fe14B1 magnets is limited, which may lead to their faster aging. Nevertheless, the permanent magnet market is currently largely monopolized by RE-based alloys, namely Nd-Fe-B magnets and Sm-Co magnets for high-temperature operation [16,17,18].
However, all RE elements are considered critical materials, and their use in wind generators, electric vehicles, and other applications that require powerful magnets is often minimized. Because some RE elements are not naturally abundant, numerous studies have been committed to the creation of high-performance permanent magnets with a diminished content of the critical RE materials. The Nd content of the Nd2Fe14B1 compound is 11.8 at.%. Aside from the Nd2Fe14B1 compound, the RE(Fe,Co)11M1 compound, which crystalizes with the ThMn12-type structure, has been recognized as a possible contender for high-performance permanent magnets where the content of 3d metals produces very large µ0MS and Tc, and a considerable a/c ratio results in a significant magnetocrystalline anisotropy energy (MAE) [16,17,18,19,20]. RE(Fe,Co)11M1 alloys have a decreased RE content (7.7 at.%) as compared with Nd2Fe14B1 magnets. However, the SmFe12 compound is thermodynamically unstable as a bulk material. In order to stabilize the REFe12 phase in bulk form, the replacement of Fe with a stabilizer M, where M = Ti [15,16,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92], V [27,29,30,32,37,39,40,41,44,49,82,85,90,93,94,95,96,97,98,99,100,101,102,103,104], Al [32,49,105,106], Si [27,30,32,35,49,107,108,109,110,111,112,113,114,115], Ga [40,82,85], Cr [27,29,32,37,49,92], Mn [37,92], Mo [27,29,32,37,39,49,72,106,116,117,118,119,120], Nb [121], W [27,29,32,49], Re [49], H [19,56,57,58,59,60,61,62,67,77], N [11,43,59,61,65,68,72,77,92,93,101,122,123,124], and C [59], has been studied. The composition range x for stabilizing the REFe12−xMx phase differs depending on M [125].
So far, Ti is known to be the favorite stabilizing element for REFe12−xMx alloys with x ~ 1 [49]. The samarium-based compound SmFe11Ti1 has the following magnetic properties: µ0Ha = 8.3–10.5 T, Tc = 584 K, and μ0Ms = 1.14–1.22 T [23,34], and is smaller than that of Nd2Fe14B1 (μ0Ms = 1.61 T) [4,12,23,34]. The iron-rich REFe12 compounds with the ThMn12-type structure earned attention for prospective permanent magnet utilization with regard to maximizing μ0Ms while minimizing Ti inclusion [20,31,35,38,46,47,52,65,68,69,71,72,73,78,84,87,88,89,90,91,96,108,109,110,111,113,115,122,126,127,128,129].
Commonly, large transition metals, such as TM = Ti, V, Cr, Nb, Mn, Mo, W, and Re, preferentially fill the 8i sites to stabilize the SmFe11TM structure [14,49,76]. Al, Si, and Ga also stabilize the structure by filling the 8f sites [49,59,76]. According to Kobayashi et al. [71], stabilization of RE(Fe,Ti)12 structures, when the Ti atoms substitute part of the smaller Fe in the 8i sites, can be considered as a reduction in the local mismatches in interatomic distances limiting the orbital hybridization. The average Fe-Fe interatomic distances for Fe (8i), Fe (8j), and Fe (8f) sites are 0.271, 0.259, and 0.251 nm, respectively, where the interatomic distance on the 8i site is longer than the Fe-Fe interatomic distance in the bcc structure (rFe = 0.126 nm, 2rFe = 0.252 nm) [71]. The replacement of some Fe atoms with atoms such as Ti (rTi = 0.147 nm) stabilizes the structure while diminishing the magnetization of Ti-free compounds [49]. Therefore, it is essential to keep the concentration of stabilizers as low as possible. The prejudiced occupancy of Fe and Ti atoms at the 8i site was also evident from a neutron diffraction experiment for the YFe11Ti1 compound [33].
According to Tozman et al. [75], the amount of magnetization of SmFe11M1 can be increased using two methods. The first method to improve the magnetization is Co replacement on the Fe site. Magnetic and structural properties of Sm(Fe1−xCox)11Ti1, Sm(Fe1−xCox)10Ti2, and Sm(Fe1−xCox)10Si2 alloys have been explored by Cheng et al. [31] and Solzi et al. [35]. It was found that the easy magnetization direction (EMD) in these alloys occurs to change from axial, SmFe11Ti1, SmFe10Si2, SmFe10.8Ti1.2, to conical, and finally to planar with elevated Co content. The planar MAE of the SmCo11Ti1 compound was also discovered by Ohashi et al. [42] and Wang et al. [55]. A diagram representing the change in EMD with temperature for each REFe11Ti1 compound (RE = Y, Pr, Nd, Sm, Gd, Th, Dy, Ho, Er, and Tm) was presented by Kou et al. [50]. Tereshina et al. [52] presented the substitution effects of Fe by Co on magnetic properties of RE(TM,Ti)12 (RE = Y, Sm; TM = Fe, Co) single crystals. They showed that REFe1−xCoxTi1 alloys have axial EMD for x ~ ≤ 0.5 μ0Ms with a maximum at x ≈ 0.3 [52]. Co has a strong influence on the MAE of the Sm-sublattice; the first anisotropy constant, K1, displays a maximum at x ≈ 0.1 and drastically decreases at x > 0.1. Finally, the Curie temperature grows monotonically with an increase in Co content [31,35,52].
The second method to improve the magnetization is to diminish the Ti content at 8i sites. Phase stability can be reached by Zr or Y replacement for Sm, which decreases the resident substructure encompassing the Fe (8i) sites. Suzuki et al. [65] utilized Zr to develop a (Nd,Zr)(Fe,Co)11.5Ti0.5N magnet. In order to diminish the lattice mismatch in interatomic distances quoted above, the part of the RE element in the 2a sites can be substituted with the smaller Zr atoms (the atomic radius of Zr is ~12% less than that of Nd). Thus, a combined alloying with both Zr and TM (for example, Ti) could make possible to achieve stability with the ThMn12-type magnet with less than one TM (for example Ti) atom per formula unit. Kuno et al. [69] mentioned that the hard magnetic characteristics of strip-cast (Sm0.8Zr0.2)(Fe0.75Co0.25)11.5Ti0.5 alloy are μ0Ms = 1.58 T, μ0Ha = 7.41 T at low temperature (T = 20 °C), which are similar to those of the Nd2Fe14B1 magnet (µ0MS = 1.61 T, µ0Ha = 7.6 T). However, the Curie temperature (Tc = 880 K) significantly outperforms the Nd2Fe14B1 magnet (Tc = 584 K).
Hagiwara et al. [73,78] synthesized via arc melting (and annealed) the (Sm0.8Y0.2)(Fe0.8Co0.2)11.5Ti0.5 magnet with µ0MS = 1.50 T, µ0Ha = 11 T at room temperature and Tc = 820 K. Both the μ0Ha and Tc significantly outperform those of the Nd2Fe14B magnet.
Sakuma et al. 2016 [123] discovered that Zr replacement stabilizes the (Nd1−αZrα)(Fe0.75Co0.25)11.25Ti0.75N1.2−1.4 (α = 0–0.3) magnets and displays a higher μ0MS = 1.62 T at α = 0.1 than the Nd2Fe14B1 magnet. According to Tozman et al. [75], for the Sm(Fe1−xCox)11Ti1 alloys, the largest μ0Ms = 1.43 T and µ0Ha = 10.9 T at room temperature are accomplished for the Sm0.94(Fe0.81Co0.19)11Ti1.08 magnet (x = 0.19), with a corresponding Curie temperature of 800 K. The μ0Ms could be improved by decreasing the Ti replacement for Fe from Ti1 to Ti0.5 and with partial replacement of Sm by Zr. According to Tozman et al. [75], for the (Sm0.77Zr0.24)(Fe0.80Co0.19)11.5Ti0.65 magnet, μ0Ms = 1.53 T and µ0Ha = 8.4 T, with Tc = 830 K.
The REFe12 compounds are unstable as bulk materials, although they have been seen in thin films [11,13,75,80,82,93,124,126,127,128,129,130,131]. Hirayama et al. [126] synthesized epitaxial Sm(Fe1−xCox)12 films (x = 0, 0.1 and 0.2) (without stabilizing elements) on W- and V-buffered single-crystalline MgO substrates. The intrinsic hard magnetic properties of the Sm(Fe0.8Co0.2)12 thin films are higher than those of Nd2Fe14B1 at room temperature (μ0Ms = 1.78 T, μ0HA = 12 T, (BH)max = 630 kJ/m3), while the Curie temperature is also improved (TC = 859 K). Additionally, Tozman et al. [80,84] reported that μ0Ms increases via partial replacement of Sm in Sm(Fe0.8Co0.2)12 with Zr. For the (Sm0.74Zr0.26)(Fe0.8Co0.2)12 magnet, μ0Ms = 1.86 T, µ0Ha = 9.8 T, and Tc = 675 K [80,84]. Ogawa et al. [129] evaluated the magnetic hardness factor, κ = (K1/μ0Ms2)1/2 of the Sm(Fe1−xCox)12 magnets at 300 K and found that κ is equal to 1.38, 1.28, and 1.24 for x = 0.0, 0.07, and 0.2, respectively. Because κ should be larger than one for a compound to be regarded as a permanent magnet material [18], the Sm(Fe1−xCox)12 alloys have a prospective use as the main phase of the high-performance permanent magnet material if the ThMn12-type phase can be bulk-stabilized.
Table 1 summarizes the intrinsic properties of some ThMn12-type structure magnets together with properties for Neomax (Nd2Fe14B1). Note that the Sm(Fe0.8Co0.2)12 and (Sm0.74Zr0.26)(Fe0.8Co0.2)12 magnets are epitaxially grown thin films and are unstable in bulk.
Multiple theoretical approaches have investigated RE(Fe,Co)11TM1 magnets. These include augmented spherical waves within the local spin density-functional (ASW-LSDF), molecular dynamics (MD) and Monte Carlo (MC), embedded atom method (EAM), full-potential linearized muffin orbital (FPLMTO), tight-binding linear-muffin orbital atomic sphere approximation (TB-LMTO-ASA) in conjunction with computational high-throughput screening (HTS), hybrid-exchange density functional theory (HE-DFT), linear combination of atomic orbitals (LCAO), pseudopotential Vienna ab initio (VASP) package, full-potential linearized augmented plane wave (FLAPW) within WIEN2K package, Hubbard I method (LDA + U), Quantum Materials Simulator and projected augmented-wave (QMAS-PAW), Korringa–Kohn–Rostoker coherent-potential approximation (AkaiKKR-CPA), mean field approximation (MFA), and Bayesian optimization (BO) approaches [11,85,92,125,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148].
Coehoorn [125] performed ASW-LSDF calculations for the YFe12 and YFe8TM4 (TM = Ti, V, Cr, Mn, Mo, and W) compounds with the ThMn12-type structure, in which the TM atoms occupy the (8i) sites. These calculations showed that the YFe12 compound is thermodynamically unstable. The calculated reduction in saturation magnetization of the YFe12−xMx (1 ≤ x ≤ 3) compounds after substituting Fe(8i) atoms by an M atom is in a good accordance with the experimental data [125]. Chen et al. [132] used classical MD to show that each TM = (Cr, V, Mo, and Ti) substantially reduces the cohesive energy of the Sm(Fe,TM)12 compound and stabilizes the ThMn12-type structure. The sequence of site preference occupation is 8i > 8j > 8f, with the 8i occupation resulting in extreme energy reduction.
Ke and Johnson [133] used the FPLMTO method to study the intrinsic properties of the RE(Fe1−xCox)11TiZ alloys, where RE = Y and Ce, interstitially dopped by Z = H, C, and N. The calculated μ0Ms, μ0Ha, and Tc values of the YFe11Ti1, YCo11Ti1, CeFe11Ti1, and CeCo11Ti1 compounds agree with the experimental data, although the Ti atom should have a strong preference to occupy the 8i sites, as observed by Moze et al. [33]. The site inclination of Co atoms in the Y(Fe1−xCox)11Ti1 compound with x < 0.4 is: 8j > 8f >8i concurs with the neutron scattering experiments by Liang et al. [131]. According to Ke and Johnson [133], the maximum magnetization of the Y(Fe1−xCox)11Ti1 and Ce(Fe1−xCox)11Ti1 alloys occurs at x = 0.2, while in experiments it takes place at x = 0.3 [54] and x = 0.1–0.15 [15] for the Y(Fe1−xCox)11Ti1 and Ce(Fe1−xCox)11Ti1 alloys, respectively. In addition, the calculated Tc of the RECo11Ti1 compound is significantly higher than the calculated Tc of the REFe11Ti1 compound [133], which agrees well with the experimental data [64].
Körner et al. [134] performed theoretical investigations (HTS-TB-LMTO-ASA) of the 1280 compounds derived from the ThMn12-type crystal structure in order to estimate their μ0Ha and (BH)max. HTS has led to several auspicious phases such as NdFe12X or NdFe11TiX (X = B, C, N) with (BH)max~600 kJ/m3 and µ0Ha~10 T. Butcher et al. [135] used the HTS-TB-LMTO-ASA methodology to study the influence of the substituted atom site preference on μ0Ms and μ0Ha of the REFe11A-type compound, where A = Al, Si, Ti, V, Cr, Co, Ni, and Cu. Körner et al. [136] used the HTS-TB-LMTO-ASA methodology to screen the magnetic properties of over 2000 different RE-TM-X compounds with 1-11 and 1-11-X (the YNi9In2-type) structures as prospective candidates for new hard-magnetic phases.
Skelland et al. [137] used the combined MD, Morse potentials, and modified EAM (MEAM) to investigate the probability distributions for Ti substitutions in RETM12−xTix (RE = Nd and Sm; TM = Fe and Co) alloys over the atom positions within the structure. The authors concluded that Ti substitution, while stabilizing the structure, also results in a considerable reduction in the spin magnetic moment due to Ti aligning anti-parallel to the net internal magnetization and reducing the magnetic moment of the neighboring iron atoms. The 8i position has the best probability for substitution followed by the 8j and 8f positions, and this conclusion agrees with the results by Miyake et al. [138].
Martinez-Casado et al. [79] performed an ab initio study of the electronic structure of CeFe11Ti1. The authors used the HE-DFT methodology, which incorporates a combined approach of atom-centered LCAO, PW-VASP, and FPLMTO. The calculations confirmed that Ti stabilizes the CeFe11Ti structure and Ce+3 causes the large MAE. Odkhuu et al. [139] performed ab initio (Wein2k, FP-LDA + U, VASP) calculations of μ0Ms and μ0Ha of the SmFe12 compound dopped by B, C, and N, indicating a possible way to improve these properties through interstitial doping of the SmFe12 compound by 2p electron elements.
Hirayama et al. and Miyake et al. [11,138] (QMAS-PAW) found that Ti substitution in the NdFe12 compound diminishes μ0Ms substantially, whereas it scarcely alters MAE. They also showed that interstitial nitrogenation substantially raises μ0Ms and the uniaxial MAE of the NdFe11Ti compound. Using the same QMAS-PAW methodology, Harashima et al. [140] proposed that in the NdFe11TiX compound, where X = B, C, N, O, and F, nitrogen is the most suitable dopant for permanent magnets in terms of μ0Ms and MAE. Harashima et al. [92] performed first-principles calculations (QMAS-PAW) of the magnetic properties of the NdFe11M compound, where M = Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn. They discovered that the calculated formation energy of the NdFe11Co compound is negative compared to the Nd2Fe17 compound (the Th2Zn17-type structure), bcc-Fe, hcp-Co, and the CoFe compound (CsCl-type structure), indicating that Co could stabilize the ThMn12 structure in bulk form. Contrary to Ti replacement, Co replacement does not diminish μ0Ms meaningfully. Harashima et al. [141] studied (QMAS-PAW) the effects of interstitial nitrogenation on the magnetic properties of the ThMn12-type compounds REFe11−xTix (x = 0, 1) with RE = Nd, Sm, Y. They found that the evolution in MAE by nitrogenation is self-sufficient of Ti replacement and does not seem to have much dependence on the RE element [141]. Matsumoto et al. [142] performed finite temperature ab initio (QMAS-PAW, AkaiKKR) MC calculations of μ0Ms and MAE to find improvements in the indirect 4f(Nd)-3d(Fe) exchange coupling to enhance the high-temperature properties of the NdFe12N-type compounds.
Using AkaiKKR-CPA formalism, Fukuzawa et al. [143] assessed intersite magnetic exchange couplings in the NdFe12 and NdFe12X (X = B, C, N, O, and F) compounds with the ThMn12 structure. They found that that MAE deteriorated rapidly with increasing temperature in the X = N system, albeit nitrogenation was dominant compared to other dopants in terms of improving the low-temperature magnetic properties. The MFA, with exchange couplings as input parameters, was used to calculate Tc. The introduction of X (X = B, C, N, O, and F) causes substantial variations in the magnetic couplings of the NdFe12 compound and has a momentous consequence in terms of Tc. Fukazawa et al. [144] presented a first-principles study of REFe12−xCrx (RE = Y, Nd, Sm) compounds with the ThMn12 structure using the QMAS -PAW package to calculate the lattice parameters of the REFe12−xMx (M = Cr, Co) system. In this study, Fe and the dopants (Co or Cr) were treated within the KKR-CPA (AkaiKKR method), and Tc was calculated within MFA. The calculations indicated an improvement in Tc over the addition of Cr, x ≤ 0.5 in the REFe12 REFe12−xCrx compound. On the contrary, above x > 0.5, Tc diminishes as the Cr concentration increases.
Harashima et al. [145] performed an ab initio investigation (QMAS-PAW) to study the effects of RE(M) site replacement and pressure on the structural stability of the REFe12 compounds, where RE = La, Pr, Nd, Sm, Gd, Dy, Ho, Er, Tm, Lu, Y, or Sc, and the group-IV element M = Zr or Hf. They found that the formation energy has a firm resemblance with the atomic radius of RE metal, in particular the formation energy with respect to simple substances diminishes as the atomic radius diminishes, except for RE = Sc and M = Hf. The calculated formation energies with respect to the RE2Fe17 compound (in both Th2Zn17 and Th2Ni17-type structures) and bcc Fe were also reported. These calculations show that the formation energy is positive, which means that one cannot make the REFe12 phase more stable than the RE2Fe17 phase, and partial replacement for Fe is needed to stabilize the ThMn12 structure.
Fukazawa et al. [146] used BO to optimize the composition of the REFe12-type compound; specifically, the composition of (RE1−α,Zα)(Fe1−βCoβ)12−γTiγ (RE= Nd, Sm, Y; Z = Zr, Dy) was improved in terms of the formation energetics, μ0Ms, μ0Ma and Tc. This scheme is based on the ab initio calculations (QMAS-PAW) and implementation of the KKR-CPA (AkaiKKR-CPA) technique to treat the compositional disorder. The authors identified 10 ThMn12-type systems with the highest value of μ0Ms, although only Nd(Fe0.7Co0.3)12 with μ0Ms = 1.89 T and Tc = 1143 K and (Sm0.7Zr0.3)(Fe0.9Co0.1)12 with μ0Ms = 1.77 T and Tc = 1002 K have the negative formation energy with respect to the unary system. The highest Tc = 1310 K was projected for the Sm(Fe0.2Co0.8) alloy, which has a negative formation energy in respect to the unary system and a significant μ0Ms = 1.47 T.
Dirba et al. [85], both experimentally and theoretically, studied how the particle size and alloy composition influence the thermal stability of the ThMn12-type phase. A definite stabilizing effect is accomplished by replacing Fe with elements such as Ti, V, or Ga. However, aligning to ab initio calculations (QMAS-PAW), the SmFe11M compound, where M = Ti and V, shows lower energy in respect to the combination of α-Fe and SmFe2 (resistance to decomposition), with preferable occupation of 8i sites for M. The SmFe11Ga compound (M = Ga) also exhibits lower energy with respect to the combination of α-Fe and SmFe2 with preferential occupation of 8j sites for Ga, but is not as stabilized as for M = Ti and V. However, the Ga dopant shows a smaller reduction in μ0Ms in comparison with Ti or V dopants. Therefore, Ga is the most successful stabilizing element when μ0Ms is considered.
Matsumoto et al. [147] used combined ab initio (AkaiKKR-CPA) methodology and experimental data from neutron diffraction to inspect the optimal chemical composition of the SmFe12 compound doped by Zr, Ti, and Co. They found that in the Zr-substituted SmFe12 magnet, Zr occupies both Sm(2a) and Fe(8i) roughly equally in terms of energetics. While using Ti as the stabilizing element cannot be avoided, it was found that the lower bound is around 0.5 Ti atoms, which is close to what has been attained experimentally [65,68,69,75]. The Zr dopant also improves the structural stability and μ0Ms of the SmFe12 magnet, which is in perfect agreement with the experimental data [80]. Dasmahapatra et al. [148] performed an ab initio study (LCAO-CRYSTAL and PW-CASP) on the effects of doping the CeFe11Ti compound with Co and Ni. The authors of [149] showed that the addition of Co (Ni) slightly improves (diminishes) μ0Ms and μ0Ha.
In our previous papers [149,150,151], we suggested increasing μ0Ms and (BH)max in the well-known SmCo5 and YCo5 magnets by substituting Co with Fe and using Ni as a thermodynamic stabilizer.
In most studies, the REFe12 (ThMn12-type structure) magnets have been stabilized in the form RE(Fe,Co)11M1, where M1 = Ti, V, Al, Si, Ga, Cr, Mn, Mo, Nb, Ta, W, Re, H, and N. A handful of papers have been published where the REFe12 magnet dopped with Ni was studied. Yang et al. [91] studied the crystallographic and magnetic properties of the replaced Y(Fe1−xMx)11Ti1 alloys, where M = Co, Ni, and Si. The thermomagnetic analysis indicated that μ0Ms has a maximum at x = 0.07 for the Y(Fe1−xNix)11Ti1 alloys. Tc increases when Fe is substituted by Ni. According to Yang et al. [91], the increases in μ0Ms and Tc can be assumed to be preferential site occupation by Ni atoms.
According to Harashima et al. [92], Ni substitution (NdFe11Ni1) yields negative formation energy values relative to the NdFe12 compound, bcc Fe, and fcc Ni, and the most stable site for Ni is 8j. Ni substitution does not diminish the magnetization significantly, as in the case of Ti, V, Cr, and Mn dopants.
Suski et al. [152] studied the structure and magnetic properties of the SmMn12–xNix alloys, which are stable in the ThMn12-type structure for the Ni concentration, 4 ≤ x ≤ 8. Suski et al. [152] asserted that the SmMn12–xNix alloys are magnetic at the low Néel temperature TN ≈ 60 K, due to antiferromagnetic ordering in the Sm sublattice, whereas the Mn and Ni sublattices are nonmagnetic. Suski et al. [153] studied the structure and magnetic properties of the NdMn12–xNix alloys, which are stable in the ThMn12-type structure for the Ni concentration of 3.6 ≤ x ≤ 6. The absence of any maximum in the temperature reliance of the magnetic susceptibility indicates the investigated alloys are nonmagnetic.
The main intent of the present research is to investigate the effects of nickel on the phase stability of the Sm(Fe1–xCox)12 alloys and to evaluate their magnetic properties. We utilize ab initio calculations using the following techniques: (i) the fully relativistic exact muffin-tin orbital method (FREMTO) and (ii) the full-potential linear muffin-tin orbital method (FPLMTO). Both methods include all relativistic effects, such as spin–orbit coupling (SOC). The application of these two methods allows us to establish accurate results that are independent of technical implementation and relies on the individual strength and durability of each method. Related details of the ab initio computational methods are outlined in Section 2. The results of the density-functional theory (DFT) calculations of the ground-state properties of the Sm(Fe-Co-Ni)12 alloys are outlined in Section 3. We demonstrate the results of the DFT calculations of the magnetic properties of the Sm(Fe-Co-Ni)12 alloys in Section 4. Lastly, a discussion and concluding remarks are outlined in Section 5.

2. Computational Methods

The fundamental framework for the present modeling is based on density-functional theory (DFT). In this approach, there is an unavoidable approximation for the electron exchange and correlation interactions. Here, we apply the generalized gradient approximation for this purpose. This exchange and correlation functional works extremely well for many elements and compounds in the periodic table but cannot determine the localized nature of the electrons. This is no issue for the 3d transition elements Fe, Co, and Ni that we study, while for the rare-earth element Sm, the 4f electrons are localized and require special treatment.
The 3d transition metals Fe, Co, and Ni are generally very well described within DFT, but for very good orbital moments one can include a simple parameter-free scheme that couples the orbital moments, i.e., an orbital polarization (OP) mechanism, which is otherwise not included in conventional DFT. This recipe results in excellent spin and orbital moments for Fe, Co, and Ni [154].
Samarium metal poses a more challenging problem for DFT. Here, the 4f electrons are known to be localized (atomic-like and essentially non-bonding) with magnetic properties closer to the free Sm3+ atom but with a magnetic moment largely quenched by crystal-field effects. DFT-GGA-type calculations are not able to describe the nature of these 4f states but exaggerate their itineracy or band-like character. For energetics such as the structural stability, the DFT-GGA approximation works better than one would expect, and the formation energies are reproduced via experiments for the SmCo5 (doped with Fe and Ni) very nicely [149,150]. For spectroscopic or magnetic properties, however, one must go beyond DFT-GGA to achieve an accurate theory. In the literature, DFT in connection with large intra-atomic Coulomb repulsion and a Hubbard U parameter (DFT + U) has been applied for Sm systems [155]. The more advanced dynamical mean-field theory (DMFT) has also been utilized for SmCo5 [156]. A more practical approach, which allows for accurate MAE for SmCo5, is the application of the “standard rare-earth model” (SRM), where the 4f electrons are treated as localized by being constrained to occupy core states that do not hybridize with the band states. This technique is very similar to DMFT in the Hubbard I approximation [157]. In the SRM, the magnetic-moment contribution from the Sm 4f electrons is constrained to zero, while a small induced magnetic moment still forms on the Sm atom (~0.3 μB) from other valence electrons. This is reasonable because for Sm metals the net magnetic moment is estimated to be very small due to spin and orbital moment cancelation (0.1 μB) [158].
The SRM approach, coupled with an orbital polarization extension of DFT (see the next section), proved to be an excellent model for the magnetic anisotropy energies in SmCo5 and related systems [149,150], and we adopt it here for all MAE calculations.
Here, we employ two methods (described below)—one is very well suited for alloy calculations, and one is appropriate for MAE calculations within the standard rare-earth model. The calculations we refer to as EMTO are carried out using Green’s function based on the improved screened Korringa–Kohn–Rostoker formalism, where the one-electron potential is described by the optimized overlapping muffin-tin (OOMT) potential spheres [159,160]. Inside the potential spheres, the potential is spherically symmetric, while it is constant between the spheres. The radius of the potential spheres, the spherical potential inside these spheres, and the constant value in the interstitial region are determined by minimizing (i) the difference between the exact and overlapping potentials and (ii) the inaccuracy that is caused by the overlap between the spheres. Within the EMTO formalism, the one-electron states are resolved exactly for the OOMT potentials. As a product of the EMTO calculation, one can solve the self-consistent Green’s function of the system and the complete, non-spherically symmetric charge density. Finally, the total energy is obtained from the full charge density formalism [161]. We regard 6s, 5p, 5d, and 4f states for Sm and 4s and 3d states for Co and Fe as valence states. The Kohn–Sham orbitals are expanded in terms of spdf exact muffin-tin orbitals. The completeness of the muffin-tin basis was analyzed thoroughly in [159]. The generalized gradient approximation (GGA-PBE) [162] is chosen for the electron exchange and the correlation energy functional. Integration over the Brillouin zone assumes a 25 × 25 × 25 grid of k-points applying the Monkhorst–Pack technique [163]. The moments of the density of states, which are needed for the kinetic energy and valence charge density, are calculated by integrating Green’s function over a complex energy contour with a 2.2 Ry diameter while using a Gaussian integration technique with 40 points on a semi-circle enclosing the occupied states. When included, SOC is carried out through the four-component Dirac equation [164].
Accounting for compositional disorders, the EMTO method is linked with the coherent potential approximation (CPA) [165,166]. The ground-state properties of the Sm(Fe1–x–yCoxNiy)12 alloys are acquired from EMTO-CPA calculations. The paramagnetic state of the Sm(Fe1–x–yCoxNiy)12 alloys is modeled within the disordered local moment (DLM) approximation [167,168]. The equilibrium atomic density of the alloy is attained from a Murnaghan fit to the total energy versus the atomic volume curve [169].
For the magnetic properties, we adopt the highest level of theory (least approximations) that is implemented in a full-potential scheme, i.e., no structural approximations, which includes spin–orbit interactions, as explained [170] for early and late lanthanides [171,172]. The full-potential linear muffin-tin orbital (FPLMTO) accomplishes this in ways that are detailed in [173]. For the Sm 4f electrons, we apply two fundamentally different approaches. First, as with the EMTO method, the Sm 4f electrons are treated as band electrons. In terms of magnetic approximations, we perform scalar-relativistic approximations, i.e., no spin–orbit coupling and no orbital moment (4f-band-scalar), the same approximations but including spin–orbit coupling (4f-band-SO), and finally also approximations with the orbital polarization added (4f-band-SO + OP). Of these three approaches, the 4f-band-SO is most like the EMTO method outlined above. Second, we assume the standard rare-earth model (SRM), i.e., we constrain the 4f electrons to occupy non-hybridizing core states, while all d states (on Sm and the transition metal component) are orbitally polarized (SRM-SO + OP). For the valence states, we use two energy parameters coupled with each basis function, and these parameters have different values for pseudocore states (4s and 4p) and valence states (5s, 5p, 4d, and 5f). For the transition metals, analogous set ups are used. The spin–orbit interaction and the orbital polarization operate on the d and f orbitals for best accuracy of the orbital magnetism. OP is introduced in FPLMTO in a self-consistent fashion, so there are no fixed parameters. Because of the way it is constructed, the orbital polarization often enhances spin–orbit coupling, leading to better orbital moments [173]. Both the generalized gradient and the local-density approximations (GGA-PBE and LDA) [162,174] are applied.
In computing the MAE, i.e., the difference in total energy between spins pointing in plane (100) or perpendicular to the plane along the z axis (001), only small differences are observed. To resolve these small numbers, one must be very careful in calculating the total energy for the two cases in as consistent a manner as possible. Here, we decide to remove all crystal symmetry so that the calculations are identical from a crystal symmetry standpoint. This approach is less efficient but in practice more accurate due to errors associated with differing k points and Brillouin zones, and other numerical approximations are minimized. Brillouin zone integrations around 3000 k points are carried out using Fermi–Dirac broadening corresponding to room temperature (300 K). Finally, the alloy compositions between iron and cobalt are modeled within the virtual-crystal approximation (VCA). This simply means that an Fe0.9Co0.1 alloy is treated as a one-atom species with an atomic charge of 26.1.

3. Thermodynamic Properties of the Sm(FeCo-Ni)12 Alloys

The SmFe12 compound crystallizes in the body centered orthogonal to the ThMn12-type structure (space group I4/mmm, no. 139; see Figure 1). The structure contains Sm atom in the 2a Wycoff position and 12 Fe atoms that occupy three inequivalent Wyckoff positions, 8f, 8i, and 8j, respectively. According to the Pearson symbol (tI26), the usual SmFe12 supercell contains 26 atoms but can be represented by the so-called reduced supercell with 13 atoms (1 Sm and 12 Fe) used in the present calculations.
EMTO calculations revealed that the SmFe12 compound has an equilibrium atomic volume Ω0 equal to 13.19 Å3, which is in fair accordance with previous calculations, 13.02 Å3 [141], but smaller than the experimental value for the SmFe12 films, Ω0 = 13.64 A3 [176]. The calculated formation energy of the SmFe12 compound with respect to the unary systems α-Sm and α-Fe is H = +6.13 mRy/atom, which confirms that the SmFe12 compound is thermodynamically unstable.
EMTO calculations revealed that the SmCo12 compound has an equilibrium atomic volume Ω0 = 12.42 Å3. The calculated formation energy of the SmCo12 compound in respect to the unary systems, α-Sm and α-Co, is H = +1.06 mRy/atom, which confirms that the SmCo12 compound is thermodynamically unstable.
Figure 2 shows the formation energy of Sm(Fe1−xCox)12 alloys calculated with respect to the unary systems α-Sm, α-Fe, and α-Co. We assume that Sm atoms solely occupies the 2a Wyckoff position and Fe and Co atoms are randomly distributed on 8f, 8i, and 8j Wyckoff positions (see Figure 1). We could have expected the formation energy to decrease linearly as Co atoms gradually replace Fe atoms on their sites. However, this Figure shows that 20% of the Co solute decreases the formation energy of the Sm(Fe0.8Co0.2)12 compound by about ~36% relative to the undoped SmFe12 compound. In the compositional range between Sm(Fe.5Co.5)12 and SmCo12, the formation energy changes very slowly from ~+2 mRy/atom to ~+1 mRy/atom.
The calculated (EMTO) value of the formation energy of three configurations (a) SmNi(8f)4Co(8i)4Co(8j)4, (b) SmNi(8i)4Co(8f)4Co(8j)4, and (c) SmNi(8j)4Co(8f)4Co(8i)4, with respect to the unary systems α-Sm, α-Co, and α-Ni, are +0.047 mRy/atom, −0.031 mRy/atom, and −1.65 mRy/atom, respectively. Thus, assuming that 1/3 of Co atoms are replaced by Ni atoms, one can obtain the energetically favorable configuration when 4 Ni atoms occupy all 8j sites and the remaining 8 Co atoms occupy 8i and 8f sites, resulting in a stable SmNi4Co8 compound with a formation energy H = −1.65 mRy/atom. This result agrees with the conclusion of Harashima et al. [92], who found that the Ni dopant prefers to occupy 8j sites when stabilizing the NdFe11Ni1 compound.
In order to explore the stabilizing effects of the addition of nickel metal to the Sm(Co1–xFex)12 alloys, we performed EMTO-CPA calculations of the formation energy of the quasi-binary SmNi4(Fe1−xCox)8 alloys relative to the unary systems α-Sm, α-Fe, α-Co, and α-Ni, where 4 Ni atoms occupy the 8j sublattice and occupation of the 8i and 8f sublattices gradually changes from pure Fe to pure Co metals. As can be seen from Figure 3, the formation energies of the quasi-binary SmNi4(Fe1−xCox)8 alloys are negative (stable alloys) within the whole compositional range. Even if the calculated heat of formation of the unstable SmFe12 compound (+7.38 mRy/atom) is approximately seven times as large as the heat of formation of the also unstable SmCo12 compound (+1.06 mRy/atom), the heats of the formation of the end points, SmNi(8j)4Co(8i,8f)8 and SmNi(8j)4Fe(8i,8f)8 compounds, are almost equal at −1.65 mRy/atom and −1.71 mRy/atom, respectively. Consequently, the occupation of the 8j sublattice by Ni atoms plays a critical role in defining the sign and magnitude of the heat of formation of the Sm (Fe,Co,Ni)12 compound. EMTO calculations revealed that the equilibrium atomic volumes of the SmNi(8j)4Fe(8i,8f)8 and SmNi(8j)4Co(8i,8f)8 compounds are Ω0 = 12.98 Å3 and 12.40 Å3, respectively.
As we mentioned in the Introduction, Harashima et al. 2018 [145] found that the calculated formation energy of the REFe12 compound in respect to the RE2Fe17 compound and α-Fe is positive, indicating that the REFe12 compound is unstable toward the decomposition of the RE2Fe17 compound and α-Fe. The possible reactions of the REFe12 compound have been suggested in [85,92,104,145,177]. In this paper, we studied the reactions suggested by Schönhöbel et al. [104], assuming the decomposition of the Sm(Fe1−xVx)12 alloys to the Sm2Fe17 compound, α-V, and α-Fe:
H f S m F e 12 x V x = 1 26 2 E S m F e 12 x V x E S m 2 F e 17 2 x E α V 7 2 x E α F e
Schönhöbel et al. [104] performed calculations for 26 atoms (doubled) of ThMn12-type supercells. Taking into consideration that in the EMTO method the total energy is calculated in the [Ry/atom] unit, x = 4 for the SmNi4Fe8 compound, while Equation (1) transforms (in the EMTO formalism) to:
H f S m F e 12   = 1 26 26 E S m F e 12 19 E S m 2 F e 17 7   E α F e ,
H f S m F e 8 N i 4   = 1 26 26 E S m F e 8 N i 4 19 E S m 2 F e 17 8 E α N i + E α F e
H f S m C o 12   = 1 26 26 E S m C o 12 19 E S m 2 C o 17 7   E α C o ,
H f S m C o 8 N i 4   = 1 26 26 E S m C o 8 N i 4 19 E S m 2 C o 17 8 E α N i + E α C o .
Table 2 presents the decomposition energy of the reaction RE(Fe,Co,Ni)12RE2(Fe, Co)17 + TMs. As previously mentioned, the SmFe12 and SmCo12 compounds are unstable (positive heat of formation reported in Figure 2). Moreover, each of theThMn12-type compounds without Ni is unstable relative to decomposition into the Th2Zn17-type compound and TMs. On the contrary, the SmNi4Fe8 and SmNi4Fe8 compounds have negative formation energies (see Figure 3) and are also stable toward decomposition into the Th2Zn17-type compounds and TMs. These results highlight the promising phase stability behavior for the SmNi4Fe8- and SmNi4Fe8-based magnets.

4. Magnetic Properties of the Sm(FeCo-Ni)12 Alloys

The total moment of the SmFe12 compound calculated in the present study using the EMTO method (mtotal ≈ 25.7211 μB/f.u.) at the equilibrium volume (Ω0 = 13.19 Å3) is smaller compared to the one calculated using VASP (mtotal = 26.56 μB/f.u.) [141]. Taking into consideration the calculated total moment per atom ( m a t . t o t = 1.9785   μ B ) and the calculated density of the SmFe12 compound (ρ = 7.69 g/cm3), one can estimate Ms = m a t . t o t   μ B ρ N A   M S m F e 12 = 1.39 MA/m and μ0Ms = 1.75 T, where μB = 9.274 × 10−24 Am2, NA = 6.0221 × 1023 atoms/mole, and M S m F e 12 = 61.12 g/mol (i.e., the average atomic weight per atom of the SmFe12 compound). Thus, the maximum energy product of the SmFe12 compound is B H m a x = 1 4 μ 0 M s 2 = 607.98 kJ/m3, where μ0 = 4π × 10−7  k g m s e c 2 A 2 is the permeability of free space. These predictions are in good agreement with both calculated (VASP results, μ0Ms = 1.83 T) [141] and experimental (μ0Ms = 1.64 T for 0.5 μm thick SmFe12 films) results [126].
Regarding the SmCo12 compound, the calculated (EMTO) total moment is equal to 17.4368 μB/f.u. at the equilibrium volume (Ω0 = 12.42 Å3). Thus, one can estimate Ms = 1.00 MA/m, μ0Ms = 1.26 T, and |BH|max = 315.19 kJ/m3 based on the calculated total moment per atom ( m a t . t o t = 1.3413 μB) and the calculated density (ρ = 8.20 g/cm3) of the SmCo12 compound.
Considering substitution of Fe by Ni in the SmFe12 compound to form the SmNi4Fe8 compound, where 4 Ni atoms occupy the 8j sublattice and 8 Fe atoms are equally colligated on the 8i and 8f sublattices, the resulting site-projected spin (m(s)) and orbital magnetic moments (m(o)) and the total moment (mtotal ≈ 19.9349 μB/f.u.) calculated (EMTO) at the equilibrium volume (Ω0 = 12.98 Å3) are presented in Table 3. Using the calculated total moment per atom ( m a t . t o t =1.5345 μB) and the calculated density of the SmNi4Fe8 compound (ρ = 8.19 g/cm3), Ms = 1.10 M A / m ,  μ0Ms = 1.38 T, and |BH|max = 377.55 kJ/m3 can be estimated.
By progressively substituting Fe with Co from SmNi4Fe8 to SmNi4Co8, the calculated (EMTO) site-projected spin (m(s)) and orbital moments (m(o)), as well as the total moments (mtotal) of the SmNi4(Fe0.9Co0.1)8, SmNi4(Fe0.8Co0.2)8, and SmNi4Co8 compounds, where 4 Ni atoms occupy the 8j sublattice and 8 (Fe1−xCox) atoms are equally distributed on the 8i and 8f sublattices, are presented in Table 3. In more detail (summarized in Table 4), for the SmNi4(Fe0.9Co0.1)8 compound at the equilibrium volume (Ω0 = 12.92 Å3), the calculated total magnetic moment is mtotal ≈ 19.9778 μB/f.u. Taking into consideration the calculated total moment per atom ( m a t . t o t = 1.5367 μB) and the calculated density of the SmNi4(Fe0.9Co0.1)8 compound (ρ = 8.25 g/cm3), one can estimate Ms = 1.10 MA/m, μ0Ms = 1.39 T, and |BH|max = 382.27 kJ/m3. For the SmNi4(Fe0.8Co0.2)8 compound at the equilibrium volume (Ω0 = 12.89 Å3), the calculated total magnetic moment is mtotal ≈ 19.5601 μB/f.u. Taking into consideration the calculated total moment per atom ( m a t . t o t = 1.5046 μB) and the calculated density of the SmNi4(Fe0.8Co0.2)8 compound (ρ = 8.30 g/cm3), one can estimate Ms = 1.08 MA/m, μ0Ms = 1.36 T, and |BH|max = 368.16 kJ/m3. For the SmNi4Co8 compound at the equilibrium volume (Ω0 = 12.47 Å3), the calculated total magnetic moment is mtotal ≈ 13.5833 μB/f.u. Taking into consideration the calculated total moment per atom ( m a t . t o t = 1.0449 μB) and the calculated density of the SmNi4Co8 compound (ρ = 8.77 g/cm3), one can estimate Ms = 0.78   M A / m ,  μ0Ms = 0.98 T, and |BH|max = 189.70 kJ/m3.
The compositional dependences of μ0Ms and |BH|max of the SmNi4(Fe1−xCox)8 alloys for x = 0.0, 0.1, 0.2. and 1.0 are shown in Figure 4 and Figure 5, respectively. μ0Ms initially increases and reaches a maximum at x = 0.1. This result is in an agreement with the experimental results for the SmTi1Fe11−xCox and SmTi2Fe10−xCox alloys [31,35,52], where μ0Ms reaches its maximum at x ≈ 0.2 [31,35] or x ≈ 0.3 [52]. A further Co concentration increase causes a sharp decrease in μ0Ms. The calculated trend for |BH|max as a function of the Co substitution is similar to μ0Ms.
Regarding the Curie temperature (Tc), a mean-field treatment can be formulated as [178,179]:
T c = 2 3 × E t o t D L M E t o t A F   k B
where E t o t D L M   a n d   E t o t A F are the ground-state total energies of the DLM and the AF (antiferromagnetic) states, respectively, and kB is the Boltzmann constant. Thus, an estimation of Tc can be accomplished from the total energy difference between the ferromagnetic (or antiferromagnetic) and paramagnetic (DLM) states. Nevertheless, in line with [179], the diversity between the total energies can be replaced by the diversity between the effective single-particle (one atomic specie) energies, which are instantly consociated with AF and DLM states (the so-called MFA treatment). In the present work, E t o t D L M and   E t o t A F are calculated at the equilibrium volumes for DLM and AF states, respectively.
Figure 6 shows the calculated (within the EMTO-CPA technique) Tc values of the pseudo-binary SmNi4(Fe1−xCox)8 alloys, where 4 Ni atoms occupy the 8j sublattice and 8 (Fe1−xCox) atoms are shared on the 8i and 8f sublattices. The calculated Tc values are equal to 853 K, 928 K, 995 K, and 1120 K for the SmNi4Fe8, SmNi4(Fe0.9Co0.1)8, SmNi4(Fe0.8Co0.2)8, and SmNi4Co8 alloys, respectively. These values are substantially higher than the Curie temperature of the generally used Neomax (Nd2Fe14B1) magnet (588 K) [12].
In Table 4, our FPLMTO-calculated magnetic properties for the SmNi4(Fe1−xCox)8 alloys are shown. The obtained properties depend on the approximation of the problematic 4f electrons on samarium. The simplest approach is a 4f-band model that ignores spin–orbit coupling. Better models include spin–orbit (SO) and orbital polarization (OP). The best approach is to confine the 4f electrons in core states in the standard rare-earth model with OP (SRM + OP) for the d states. The largest total moments are obtained from SRM + OP because here the Sm moment (anti-parallel to the total unit-cell moment) is the smallest.
We can compare our FPLMTO magnetic moments with those of the fully relativistic EMTO method (see Table 3). The reason these models give different total magnetic moments is because they predict vastly different samarium magnetic moment. In Figure 7, we show the different results for the models.
Figure 7 shows that the SRM + OP model has the smallest Sm moment and the greatest total magnetic moment (Table 4). The 4f-band-EMTO approach is second. We believe there are technical differences in the EMTO and FPLMTO codes that give rise to the very different Sm moments, because relativistic DFT values (with spin–orbit coupling but no orbital polarization), i.e., EMTO and 4f-band-SO (FPLMTO), should in principle be close but they are not. For the FPLMTO calculations, the Sm magnetic moment depends very weakly on composition parameter x, while there is some dependence in the EMTO. We should point out that for the 3d (Fe, Co, and Ni) metals, the magnetic moments are almost the same for all models.
Next, we present our MAE results. Here, we apply the SRM + OP model for the MAE because the other approaches (4f-band) produce unrealistically large MAE due to the inappropriate treatment of the 4f electrons (not shown). Before calculating the MAE, i.e., the energy difference between the magnetic spin moment aligned in the x-y plane and along the z axis, we carefully optimize the crystal structure. In other words, both the lattice parameters a and c are optimized to give the lowest total energy of the tetragonal crystal. The differences between a and c for the iron-rich compound SmNi4(Fe1−xCox)8 (here x = 0, 0.1, and 0.2) are actually very small, while for x = 1 the total atomic volume decreases by about 4%. We present the calculated equilibrium atomic volume (Ω0) and μ0Ha (with Ha = 2K1/Ms) [17]) in Table 5.
As was mentioned in [17,180], the relation (interchange) between the magnetic anisotropy and saturation magnetization is the key criterion for a magnet to be fabricated in any possible shape with a good chance of resisting self-demagnetization. This criterion may be formulated in terms of the magnetic hardness parameter of the materials, κ = K 1 μ 0 M s 2 , where κ > 1 is an empirically required rule to fabricate a good permanent magnet. As reported in Table 5, all calculated magnetic hardness parameters of the SmNi4(Fe1−xCox)8 magnets meet the manufacturability criteria (i.e., κ > 1).
Finally, the compositional dependence of the anisotropy field (μ0Ha) of the SmNi4(Fe1−xCox)8 alloys is shown in Figure 8. We find that on the iron-rich side of the compound, adding a small amount of cobalt increases μ0Ha in a linear fashion.

5. Discussion and Conclusions

The present calculations show that the SmNi4(Fe1−xCox)8 alloys could be stable; fabricated in bulk form within the whole compositional range; and have significant μ0Ms values of 1.38–1.57 T, 1.39–1.53 T, and 1.36–1.42 T (depending on the model); Tc values of 853 K, 928 K, and 995 K; μ0Ha values of 6.09 T, 8.02 T, and 10.54 T; and κ values of 1.40, 1.62, and 1.88 for x = 0.0, 0.1, and 0.2, respectively. As mentioned in the Introduction, the economic requirement for a permanent magnet is fulfilled if μ0Ms ~≥ 1.25 T, Tc ~≥ 550 K, μ0Ha ~≥ 3.75 T, and κ > 1 [1,2,17], which means the SmNi4(Fe1−xCox)8 magnets could be suitable for industrial applications. The most promising SmNi4(Fe0.9Co0.1)8 magnet outperforms most of the intrinsic properties of the widely used Neomax magnets (Tc = 588 K and μ0Ha = 7.60 T), except for μ0Ms = 1.61 T (Neomax), resulting in |BH|max = 382.27–462.5 kJ/m3, which is 74.2–89.8% of |BH|max~515 kJ/m3 (Neomax). Stabilization of the bulk SmFe12 magnet is achieved by doping it with Ti, Co, Zr, and Y. According to [138,142,181], for early transition metals (TM = Ti-Mn), the stabilization of the SmFe11TM1 magnet is accompanied by substantial reduction in the magnetic moment. In contrast, the late transition metals (TM = Co, Ni, Cu, and Zn), when serving as stabilizers, increase the magnetic moment of the SmFe12 magnet. As was mentioned in the Introduction, Harashima et al. [92] found that Ni substitution yields a negative formation energy of the NdFe11Ni1 compound relative to the NdFe12 compound, bcc Fe, and fcc Ni. Our calculations confirm the observation made by Harashima et al. [92] that the most stable site for Ni is 8j, in contrast to Ti (the most stable site for Ti is 8i), for which the spin magnetic moment aligns anti-parallel to the internal magnetization of the SmFe12 compound and the spin moment of Ni aligns parallel to the internal net magnetization of the SmFe12 compound, thereby boosting its μ0Ms.
In accordance with our previous calculations [149,150], substituting most of Co with Fe in the SmCo5 magnet (the CaCu5-type structure) and using Ni as a thermodynamic stabilizer engenders the formation of the SmFe3CoNi magnet, which has exceptional magnetic properties, such as a high Tc = 1103 K, powerful magnetic anisotropy (K1 = 9.2 MJ/m3, about twice that of the Nd2Fe14B1 magnet at K1 = 4.9 MJ/m3), and |BH|max = 362 kJ/m3, which is about 70% of |BH|max~515 kJ/m3 (Neomax). Ni has also been used as the stabilizing element in the YCo5 magnet (the CaCu5-type structure) in order to provide the maximum amount of Fe to favor high magnetization of the YFe3CoNi magnet |BH|max~309 kJ/m3 [151].
When considering why Ni acts as a stabilizer, it is important to realize that the favored structure depends rather sensitively on the number of d electrons in the compound, i.e., the filling (occupation) of electrons in the d-band. Figure 9 shows the calculated formation energies (EMTO) of the SmFe12, SmCo12, and SmNi12 compounds in respect to the unary systems α-Sm, α-Fe, α-Co, and α-Ni as a function of the number of 3d electrons per TM atom.
Increasing the 3d electron count (more Ni) greatly stabilizes the SmTM12 compound. An obvious consequence of this behavior is that one should be able to recover the crystal stability of a Sm(Fe,Co)12 compound by doping it with 3d electrons from nickel. From Figure 9, one can conclude that an Sm(Fe1−x−yCoxNiy)12 alloy with at least 7.76 3d electrons per TM atom is stable. This result is consistent with the previously discovered stabilization of the hexagonal SmTM5 (the CaCu5-type structure) compound [149]. According to Söderlind et al. [149], the efforts to enhance the magnetism of the stable SmCo5 magnets through the substitution of some amount of Co for Fe were limited because the SmFe5 compound does not exist in the CaCu5-type structure. To compensate for the reduction in d electrons due to the Fe/Co substitution (Fe has about one 3d electron less than Co), we simply added nickel, which has about one more 3d electron than cobalt, so that the structural stability was ensured. Nickel acts as a thermodynamic stabilizer in the Sm-Fe-Co-Ni alloys, and in present case, adding a certain amount of Ni helps to stabilize the SmNi4(Fe1−xCox)8 magnets in the ThMn12-type structure.
Our calculations revealed that the SmNi4Co8 magnet exhibits a significant axial MAE with K1 = 9.41 MJ/m3. As is noted in the Introduction, the planar MAE of the SmCo11Ti compound was discovered by Ohashi et al. [42] and Wang et al. [55]. Tereshina et al. [52] concluded that Co has a strong influence on the MAE of Sm-sublattice in the case of the REFe1−xCoxTi alloys, causing changes in MAE from axial to planar when x ~≥ 0.5. Our calculations did not reproduce this behavior.
In conclusion, we showed that the Sm(Fe1−xCox)12 alloys, including the end points SmFe12 and SmCo12, could be stabilized by replacing a certain portion of Fe or Co atoms with Ni atoms and could perhaps be synthesized. These contemporary permanent magnets are assumed to possess outstanding magnetic characteristics, namely a significant energy product and anisotropy field, as well as a high Curie temperature that significantly outperform Neomax magnets. Nevertheless, it is essential to accentuate that although these intrinsic properties are required for profitable permanent magnets, this is not a plentiful condition. It is also crucial to advance the anisotropic polycrystalline microstructure of a magnet to provide both high coercivity and remanence, e.g., to develop an applicable grain boundary phase that can grow in equilibrium with the main phase, in order to optimize the extrinsic magnetic properties.

Author Contributions

Conceptualization, A.L.; methodology, A.L. and P.S.; writing-review and editing, A.L., P.S., A.P. and E.E.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the Critical Materials Institute, an Energy Innovation Hub funded by the US Department of Energy, Office of Energy Efficiency and Renewable Energy, Advanced Manufacturing Office.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The work was performed under the auspices of the US Department of Energy by the Lawrence Livermore National Laboratory under Contract No. DE-AC52-07NA27344. A.L. thanks A. Ruban, O. Peil, P. Korzhavyi, L. Vitos, and M. Däene for their technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Buschow, K.H.J. New developments in hard magnetic materials. Rep. Prog. Phys. 1991, 54, 1123–1214. [Google Scholar] [CrossRef]
  2. Buschow, K.H.J.; de Boer, F.R. Physics of Magnetism and Magnetic Materials; Kluwer Academic/Plenum: New York, NY, USA, 2004. [Google Scholar]
  3. Sagawa, M.; Fujimura, S.; Togawa, N.; Yamamoto, H.; Matsuura, Y. New material for permanent magnets on a base of Nd and Fe (invited). J. Appl. Phys. 1984, 55, 2083–2087. [Google Scholar] [CrossRef]
  4. Sagawa, M.; Fujimura, S.; Yamamoto, H.; Matsuura, Y.; Hirosawa, S. Magnetic properties of rare-earth-iron-boron permanent magnet materials. J. Appl. Phys. 1985, 57, 4094–4096. [Google Scholar] [CrossRef]
  5. Sagawa, M.; Hirosawa, S.; Yamamoto, H.; Fujimura, S.; Matsuura, Y. Nd–Fe–B Permanent Magnet Materials. Jpn. J. Appl. Phys. 1987, 26, 785–900. [Google Scholar] [CrossRef]
  6. Sagawa, M.; Hirosawa, S.; Tokuhara, K.; Yamamoto, H.; Fujimura, S.; Tsubokawa, Y.; Shimizu, R. Dependence of coercivity on the anisotropy field in the Nd2Fe14B type sintered magnets. J. Appl. Phys. 1987, 61, 3559–3561. [Google Scholar] [CrossRef]
  7. Hirosawa, S.; Matsuura, Y.; Yamamoto, H.; Fujimura, S.; Sagawa, M.; Yamauchi, H. Magnetization and magnetic anisotropy of Nd2Fe14B measured on single crystals. J. Appl. Phys. 1986, 59, 873–879. [Google Scholar] [CrossRef]
  8. Matsuura, Y. Recent development of Nd-Fe-B sintered magnets and their applications. J. Magn. Magn. Mater. 2006, 303, 344–347. [Google Scholar] [CrossRef]
  9. Sepehri-Amin, H.; Ohkubo, T.; Nishiuchi, T.; Hirosawa, S.; Hono, K. Coercivity enhancement of hydrogenation–disproportionation–desorption–recombination processed Nd–Fe–B powders by the diffusion of Nd–Cu eutectic alloys. Scr. Mater. 2010, 63, 1124–1127. [Google Scholar] [CrossRef]
  10. Akiya, T.; Liu, J.; Sepehri-Amin, H.; Ohkubo, T.; Hioki, K.; Hattori, A.; Hono, K. Low temperature diffusion process using rare earth-Cu eutectic alloys for hot-deformed Nd-Fe-B bulk magnets. J. Appl. Phys. 2014, 115, 17A766. [Google Scholar] [CrossRef]
  11. Hirayama, Y.; Miyake, T.; Hono, K. Rare-earth lean hard magnet compound NdFe12N. JOM 2015, 67, 1344–1349. [Google Scholar] [CrossRef]
  12. Hirosawa, S.; Nishino, M.; Miyashita, S. Perspectives for high-performance permanent magnets: Applications, coercivity, and new materials. Adv. Nat. Sci. Nanosci. Nanotechnol. 2017, 8, 013002-1–013002-12. [Google Scholar] [CrossRef] [Green Version]
  13. Sepehri-Amin, H.; Tamazawa, Y.; Kambayashi, M.; Saito, G.; Takahashi, Y.K.; Ogawa, D.; Ohkubo, T.; Hirosawa, S.; Doi, M.; Shima, T.; et al. Achievement of high coercivity in Sm(Fe0.8Co0.2)12 anisotropic magnetic thin film by boron doping. Acta Mater. 2020, 194, 337–342. [Google Scholar] [CrossRef]
  14. Coey, J.M.D. Perspective and prospects for rare earth permanent magnets. Eng. J. 2020, 6, 119–131. [Google Scholar] [CrossRef]
  15. Goll, D.; Loeffler, R.; Stein, R.; Pflanz, U.; Goeb, S.; Karimi, R.; Schneider, G. Temperature dependent magnetic properties and application potential of intermetallic Fe11−XCoXTiCe. Phys. Status Solidi 2014, 8, 862–865. [Google Scholar]
  16. Niarchos, D.; Gjoka, M.; Schönhöbel, A.M.; Aubert, A.; Madugundo, R.; Garitaonandí, J.J.S.J.; Barandiaranb, J.M.; Hadjipanayi, G. Intrinsic magnetic properties of (Nd1−xSmx)Fe11Ti. J. Alloys Compd. 2021, 864, 158097. [Google Scholar] [CrossRef]
  17. Coey, J.M.D. Hard magnetic materials: A Perspective. IEEE Trans. Magn. 2011, 47, 4671–4681. [Google Scholar] [CrossRef]
  18. Coey, J.M.D. Permanent magnets: Plugging the gap. Scr. Mater. 2012, 67, 524–529. [Google Scholar] [CrossRef]
  19. Isnard, O.; Guillot, M.; Miraglia, S.; Fruchart, D. High field magnetization measurements of SmFe11Ti and SmFe11TiH1−d. J. Appl. Phys. 1996, 79, 5542–5544. [Google Scholar] [CrossRef]
  20. Bessais, L.; Djega-Mariadassou, C. Structure and magnetic properties of nano-crystalline Sm(Fe1xCox)11Ti (x ≤ 2). Phys. Rev. B 2001, 63, 54412. [Google Scholar] [CrossRef]
  21. Hadjipanayi, G.C.; Aly, S.H.; Chang, S.-F. Hard magnetic properties of R-Fe-Ti. Appl. Phys. Lett. 1987, 51, 2048–2050. [Google Scholar] [CrossRef]
  22. Ohashi, K.; Yokoyama, T.; Osugi, R.; Tawara, Y. The magnetic and structural properties of R-Ti-Fe ternary compound. IEEE Trans. Magn. 1987, 23, 3101–3103. [Google Scholar] [CrossRef]
  23. Ohashi, K.; Tawara, Y.; Osugi, R.; Shimao, M. Magnetic properties of Fe-rich rare-earth intermetallic compounds with a ThMn12 structure. J. Appl. Phys. 1988, 64, 5714–5716. [Google Scholar] [CrossRef]
  24. Ohashi, K.; Tawara, Y.; Osugi, R.; Sakurai, J.; Komura, Y. Identification of the intermetallic compound consisting of Sm, Ti, Fe. J. Less-Common. Met. 1988, 139, L1–L5. [Google Scholar] [CrossRef]
  25. Liu, N.C.; Kamprath, N.; Wickramasekara, L.; Cadieu, F.J.; Stadelmaier, H.H. Crystal Structure of R(Ti,Fe)12 (R = Nd, Sm) compound. J. Appl. Phys. 1988, 63, 3589–3591. [Google Scholar] [CrossRef]
  26. Hu, J.; Wang, T.; Zhang, W.S.; Wang, Y.; Wang, Z. Structure and magnetic properties of RTi1.1Fe10.9. J. Mag. Mag. Mater. 1988, 74, 22–26. [Google Scholar] [CrossRef]
  27. Buschow, K.H.J. Structure and properties of some novel ternary Fe-rich rare-earth intermetallics (invited). J. Appl. Phys. 1988, 64, 3130–3135. [Google Scholar] [CrossRef]
  28. Yang, Y.-C.; Hong, L.-H.; Sun, S.-H.; Gu, D.-M. Intrinsic magnetic properties of SmTiFe10. J. Appl. Phys. 1988, 63, 3702–3703. [Google Scholar] [CrossRef]
  29. Verhoef, R.; de Boer, F.; Zhi-dong, Z.; Buschow, K. Moment reduction in RFe12−xTx compounds (R = Gd, Y and T-Ti, Cr, V, Mo, W). J. Magn. Magn. Mater. 1988, 75, 319–322. [Google Scholar] [CrossRef]
  30. Solzi, M.; Pareti, L.; Moze, O.; David, W.I.F. Magnetic anisotropy and crystal structure of intermetallic compounds of the ThMn12 structure. J. Appl. Phys. 1988, 54, 5084–5087. [Google Scholar] [CrossRef]
  31. Cheng, S.F.; Sinha, V.K.; Xu, Y.; Elbicki, J.M.; Boltich, E.B.; Wallace, W.E.; Sankar, S.G.; Laughlin, D.E. Magnetic and structural properties of SmTiFe11−xCox alloys. J. Magn. Magn. Mater. 1988, 75, 330–338. [Google Scholar] [CrossRef]
  32. De Mooij, D.B.; Buschow, K.H.J. Some novel ternary ThMn12-type compounds. J. Less-Common. Met. 1988, 136, 207–215. [Google Scholar] [CrossRef]
  33. Moze, O.; Pareti, L.; Solzi, M.; David, W.I.F. Neutron diffraction and magnetic anisotropy study of Y-Fe-Ti intermetallic compounds. Solid State Commun. 1988, 66, 465–469. [Google Scholar] [CrossRef]
  34. Hu, B.P.; Li, H.S.; Gavigan, J.P.; Coey, J.M.D. Intrinsic magnetic properties of the iron-rich ThMn12-sinvitedtructure alloys R(Fe11Ti); R = Y, Nd, Sm, Gd, Tb, Dy, Ho, Er, Tm and Lu. J. Phys. Condens. Matter 1989, 1, 755–770. [Google Scholar] [CrossRef]
  35. Solzi, M.; Xue, R.H.; Pareti, L. Magnetic anisotropy and first-order magnetization processes in Sm(Fe1−xCox)10M2 (M = Ti, Si) compounds. J. Magn. Magn. Mater. 1990, 88, 44–50. [Google Scholar] [CrossRef]
  36. Hu, B.P.; Li, H.S.; Coey, J.M.D. Relationship between ThMn12 and Th2Ni17 structure types in the YFe11−xTix alloy series. J. Appl. Phys. 1990, 87, 4838–4840. [Google Scholar] [CrossRef]
  37. Okada, M.; Kojima, A.; Yamagishi, K.; Homma, M. High coercivity in melt-spin SmFe10(Ti,M)2 ribbons (M = V/Cr/Mn/Mo). IEEE Trans. Magn. 1990, 26, 1376–1378. [Google Scholar] [CrossRef]
  38. Wecker, J.; Katter, M.; Schnitzke, K.; Schultz, L. Magnetic hardening of Sm-Fe-Ti alloys. J. Appl. Phys. 1990, 67, 4951–4953. [Google Scholar] [CrossRef]
  39. Schultz, L.; Schnitzke, K.; Wecker, J. Magnetic hardening of Sm-Fe-Mo, Sm-Fe-V, Sm-Fe-Ti. J. Magn. Magn. Mater. 1990, 83, 254–256. [Google Scholar] [CrossRef]
  40. Wang, Y.; Hadjipanayis, G.C.; Kim, A.; Li, N.C.; Sellmyer, D.J. Magnetic and structure studies of Sm-Fe-Ti. J. Appl. Phys. 1990, 67, 4954–4956. [Google Scholar] [CrossRef]
  41. Wang, Y.; Hadjipanayis, G.C. Magnetic properties of Sm-Fe-Ti-V alloys. J. Magn. Magn. Mater. 1990, 87, 375–378. [Google Scholar] [CrossRef]
  42. Ohashi, K.; Ido, H.; Konno, K.; Yoneda, Y. Magnetic properties of RTiCo11 compounds with the ThMn12 structure. J. Appl. Phys. 1991, 70, 5986–5988. [Google Scholar] [CrossRef]
  43. Cadieu, F.J.; Hedge, H.; Navarathna, A.; Rani, R.; Chen, K. High-energy product ThMn12 Sm-Fe-N and Am-Fe permanent magnets synthesized as oriented sputtered films. Appl. Phys. Lett. 1991, 59, 875–877. [Google Scholar] [CrossRef]
  44. Cadieu, F.J.; Kamprath, N.; Hegde, H.; Chen, K.; Navarathna, A.; Rani, R. Analogous Sm-Ti-Fe high-coercivity phases in Sm-V-Fe and Sm-Zr-Fe sputtered film samples. J. Appl. Phys. 1991, 69, 5608–5610. [Google Scholar] [CrossRef]
  45. Cheng, S.F.; Sinha, V.K.; Ma, B.M.; Sankar, S.G.; Wallace, W.E. Phase analysis and magnetic properties of RTiFe1−xCox (R = Y, Gd) (x = 0–11). J. Appl. Phys. 1991, 69, 5605–5607. [Google Scholar] [CrossRef]
  46. Sugimoto, S.; Kojima, A.; Okada, M.; Homma, M. Enhancement of magnetic properties of Sm(Fe,Co,Ti)12 melt-spun ribbons by refining crystallized grains. Mater. Trans. JIM 1991, 32, 1180–1183. [Google Scholar] [CrossRef] [Green Version]
  47. Reinsch, B.; Grieb, B.; Henig, E.-T.; Petzow, G. Phase relations in the system Sm-Fe-Ti and the consequences for the production of permanent magnets. IEEE Trans. Magn. 1992, 28, 2832–2834. [Google Scholar] [CrossRef]
  48. Kim, H.T.; Kim, Y.B.; Kim, C.S.; Rim, T.K.; Jin, H. Magnetocrystalline anisotropy of (Sm0.5RE0.5)Fe11Ti compounds (RE = Ce, Pr, Nd, Sm, Gd and Tb). J. Magn. Magn. Mater. 1996, 152, 387–390. [Google Scholar] [CrossRef]
  49. Buschow, K.H.J. Permanent magnet materials based on tetragonal rare earth compounds of the type RFe12xMx. J. Magn. Magn Mater. 1991, 100, 79–89. [Google Scholar] [CrossRef]
  50. Kou, X.C.; Zhao, T.S.; Grössinger, R.; Kirchmayr, H.R.; Li, X.; de Boer, F.R. Magnetic phase transitions, magnetocrystalline anisotropy, and crystal-field interactions in the RFe11Ti series (where R = Y, Pr, Nd, Sm, Gd, Tb, Dy, Ho, Er, or Tm). Phys. Rev. B 1993, 47, 3231–3242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Bodriakov, V.Y.; Ivanova, T.I.; Nikitin, S.A.; Tereshina, I.S. Magnetic anisotropy and magnetoelastic properties of SmFe11Ti. J. Alloys Compd. 1997, 259, 265–269. [Google Scholar] [CrossRef]
  52. Tereshina, I.S.; Nikitin, S.A.; Ivanova, I.; Skokov, K.P. Rare-earth and transition sublattice contributions to magnetization and magnetic anisotropy of R(TM,Ti)12 single crystals. J. Alloys Compd. 1998, 275–277, 625–628. [Google Scholar] [CrossRef]
  53. Nikitina, S.A.; Tereshina, I.S.; Verbetsky, V.N.; Salamov, A.A. Transformations of magnetic phase diagram as a result of insertion of hydrogen and nitrogen atoms in crystalline lattice of RFe11Ti compounds. J. Alloys Compd. 2001, 316, 46–50. [Google Scholar] [CrossRef]
  54. Wang, J.L.; Tang, N.; Fuquan, B.; Wang, W.H.; Wang, W.Q.; Wu, G.H.; Yang, F.M. A study of the magnetocrystalline anisotropy of RFe11−xCoxTi compounds with R = Y and Er. J. Phys Condens. Matter 2001, 13, 1617–1626. [Google Scholar] [CrossRef]
  55. Wang, W.Q.; Wang, J.L.; Tang, N.; Fuquan, B.; Wu, G.H.; Yang, F.M.; Jim, H.M. Structural and magnetic properties of RCo12−xTix (R = Y and Sm) and YFe12xTix. J. Phys. D Appl. Phys. 2001, 34, 307–312. [Google Scholar] [CrossRef]
  56. Nikitin, S.A.; Tereshina, I.S.; Verbetsky, V.N.; Salamova, A.A.; Skokov, K.P.; Pankratov, N.Y.; Skourski, Y.V.; Tristan, N.V.; Zubenko, V.V.; Tereshina, I.V. Magnetostriction and magnetic anisotropy in TbFe11TiHX (x = 0, 1) single crystals. J. Alloys Compd. 2001, 322, 42–44. [Google Scholar] [CrossRef]
  57. Tereshina, I.S.; Gaczyńsky, P.; Rusakov, V.S.; Drulis, H.; Nikitin, S.A.; Suzki, W.; Tristan, N.V.; Palewski, T. Anisotropy and Mossbauer Effect Studies of YFe11Ti and YFe11TiH. J. Phys. Cond. Matter 2001, 13, 8161–8170. [Google Scholar] [CrossRef]
  58. Nikitin, S.A.; Tereshina, I.S.; Pankratov, N.Y.; Skorski, Y.V. Spin reorientation and crystal field in the single-crystal hydride HoFe11TiH. Phys. Rev. B 2001, 63, 134420. [Google Scholar] [CrossRef]
  59. Bacmann, M.; Baudelet, C.; Fruchart, D.; Gignoux, D.; Hill, E.K.; Krill, G.; Morales, M.; Vert, R.; Wolfers, P. Exchange interactions and magneto-crystalline anisotropy in RFe12−xMx and parent interstitial compounds. J. Alloys Compd. 2004, 383, 166–172. [Google Scholar] [CrossRef]
  60. Piquer, C.; Grandjean, F.; Isnard, O.; Pop, V.; Long, G.J. A magnetic and Mössbauer spectral study of the spin reorientation in NdFe11Ti and NdFe11TiH. J. Appl. Phys. 2004, 95, 6308–6316. [Google Scholar] [CrossRef]
  61. Tereshina, I.S.; Pankratov, N.Y.; Tereshina, E.A.; Verbetski, V.N.; Salamova, A.A.; Mydlarz, T.; Suski, W. Magnetocrystalline anisotropy of hydrogenated and nitrogenated rare-earth intermetallic compound SmFe11Ti. Phys. Met. Metallogr. 2005, 99 (Suppl. S1), S46–S49. [Google Scholar]
  62. Pankratov, N.Y.; Nikitin, S.A.; Iwasieczko, W.; Drulis, H.; Nenkov, K.; Skokov, O.; Gutfleisch, K.P.; Handstein, A.; Müller, K.-H. Effect of hydrogen insertion on the magnetic properties of Er(Fe,Co)11Ti single crystals. J. Alloys Compd. 2005, 404–406, 181–184. [Google Scholar] [CrossRef]
  63. Gjoka, M.; Sarafidis, C.; Kalogirou, O.; Devlin, E.; Niarchos, D. Structural and magnetic properties of Sm3(Fe1−xCox)27.7Ti1.3. IEEE Trans. Magn. 2006, 42, 3767–3769. [Google Scholar] [CrossRef]
  64. Zhou, C.; Pinkerton, F.E.; Herbst, J.F. Magnetic properties of CeFe11−xCoxTi with ThMn12 structure. J. Appl. Phys. 2014, 115, 17C716. [Google Scholar]
  65. Suzuki, S.; Kuno, T.; Urushibata, K.; Kobayashi, K.; Sakuma, N.; Washio, K.; Kishimoto, H.; Kato, A.; Manabe, A. A (Nd, Zr)(Fe, Co)11.5Ti0.5Nx compound as a permanent magnet material. AIP Adv. 2014, 4, 117131. [Google Scholar] [CrossRef]
  66. Zhou, C.; Sun, K.; Pinkerton, F.E.; Kramer, M.J. Magnetic hardening of Ce1+xFe11yCoyTi with ThMn12 structure by melt spinning. J. Appl. Phys. 2015, 117, 17A741. [Google Scholar] [CrossRef]
  67. Kostyuchenko, N.V.; Zvezdin, A.K.; Tereshina, E.A.; Skourski, Y.; Doerr, M.; Drulis, H.; Pelevin, I.A.; Tereshina, I.S. High-field magnetic behavior and forced-ferromagnetic state in an ErFe11TiH single crystal. Phys. Rev. B 2015, 92, 104423. [Google Scholar] [CrossRef]
  68. Suzuki, S.; Kuno, T.; Urushibata, K.; Kobayashi, K.; Sakuma, N.; Washio, K.; Yano, M.; Kato, A.; Manabe, A. A new magnet material with ThMn12 structure: (Nd1−xZrx)(Fe1−yCoy)11+zTi1−zNa (a = 0.6–1.3). J. Magn. Magn. Mater. 2016, 401, 259–268. [Google Scholar] [CrossRef]
  69. Kuno, T.; Suzuki, S.; Urushibata, K.; Kobayashi, K.; Sakuma, N.; Yano, M.; Kato, A.; Manabe, A. (Sm,Zr)(Fe,Co)11.0–11.5Ti1.0–0.5 compounds as new permanent magnet materials. AIP Adv. 2016, 6, 025221. [Google Scholar] [CrossRef] [Green Version]
  70. Wuest, H.; Bommer, L.; Huber, A.M.; Goll, D.; Weissgaerber, T.; Kieback, B. Preparation of nanocrystalline Ce1−xSmx(Fe,Co)11Ti by melt spinning and mechanical alloying. J. Magn. Magn. Mater. 2017, 428, 194–197. [Google Scholar] [CrossRef]
  71. Kobayashi, K.; Suzuki, S.; Kuno, T.; Urushibata, K.; Sakuma, N.; Yano, M.; Shouji, T.; Kato, A.; Manabe, A. The stability of newly developed (R,Zr)(Fe,Co)12xTix alloys for permanent magnets. J. Alloys Compd. 2017, 694, 914–920. [Google Scholar] [CrossRef]
  72. Kobayashi, K.; Furusawa, D.; Suzuki, S.; Kuno, T.; Urushibata, K.; Sakuma, N.; Yano, M.; Shoji, T.; Kato, A.; Manabe, A.; et al. High-temperature stability of ThMn12 magnet materials. Mater. Trans. 2018, 59, 1845–1853. [Google Scholar] [CrossRef] [Green Version]
  73. Hagiwara, M.; Sanada, N.; Sakurada, S. Effect of Y substitution on the structural and magnetic properties of Sm(Fe0.8.Co0.2)11.4Ti0.6. J. Magn. Magn Mater. 2018, 465, 554–558. [Google Scholar] [CrossRef]
  74. Martin-Cid, A.; Salazar, D.; Schönhöbel, A.M.; Garitaonandia, J.S.; Barandiaran, J.M.; Hadjipanayis, G.C. Magnetic properties and phase stability of tetragonal Ce1−xSmxFe9Co2Ti 1:12 phase for permanent magnets. J. Alloys Compd. 2018, 749, 640–644. [Google Scholar] [CrossRef]
  75. Tozman, P.; Sepehri-Amin, H.; Takahashi, Y.K.; Hirosawa, S.; Hono, K. Intrinsic magnetic properties of Sm(Fe1−xCox)11Ti and Zr-substituted Sm1−yZrFe0.8Co0.2)11.5Ti0.5 compounds with ThMn12 structure toward the development of permanent magnets. Acta Mater. 2018, 153, 354–363. [Google Scholar] [CrossRef]
  76. Gabay, A.; Hadjipanayis, G. Recent developments in RFe12-type compounds for permanent magnets. Scr. Mater. 2018, 154, 284–288. [Google Scholar] [CrossRef]
  77. Tereshina, I.S.; Kostyuchenko, N.V.; Tereshina-Chitrova, E.A.; Skourski, Y.; Doerr, M.; Pelevin, I.A.; Zvezdin, A.K.; Paukov, M.; Havela, L.; Drulis, H. ThMn12-type phases for magnets with low rare-earth content: Crystal-field analysis of the full magnetization process. Sci. Rep. 2018, 8, 3595–1–6. [Google Scholar] [CrossRef]
  78. Hagiwara, M.; Sanada, N.; Sakurada, S. Structure and magnetic properties of rapidly quenched (Sm,R)(Fe,Co)11.4Ti0.6. AIP Adv. 2019, 9, 035036. [Google Scholar] [CrossRef] [Green Version]
  79. Martinez-Casado, R.; Dasmahapatra, A.; Sgroi, M.F.; Romero-Muñiz, C.; Herper, H.C.; Vekilova, O.Y.; Ferrari, A.M.; Pullini, D.; Desmarais, J.; Maschio, L. The CeFe11Ti permanent magnet: A closer look at the microstructure of the compound. J. Phys. Condens. Matter 2019, 31, 505505. [Google Scholar] [CrossRef]
  80. Tozman, P.; Takahashi, Y.K.; Sepehri-Amin, H.; Ogawa, D.; Hirosawa, S.; Hono, K. The effect of Zr substitution on saturation magnetization in (Sm1−xZrx)(Fe0.8Co0.2)12 compound with the ThMn12 structure. Acta Mater. 2019, 178, 114–121. [Google Scholar] [CrossRef]
  81. Dirba, I.; Li, J.; Sepehri-Amin, H.; Ohkubo, T.; Schrefl, T.; Hono, K. Anisotropic, single-crystalline SmFe12-based microparticles with high roundness fabricated by jet-milling. J. Alloys Compd. 2019, 804, 155–162. [Google Scholar] [CrossRef]
  82. Dirba, I.; Sepehri-Amin, H.; Ohkubo, T.; Hono, K. Development of ultra-fine grain sized SmFe12-based powders using hydrogenation disproportionation desorption recombination process. Acta Mater. 2019, 165, 373–380. [Google Scholar] [CrossRef]
  83. Saito, T.; Watanabe, F.; Nishio-Hamane, D. Magnetic properties of SmFe12-based magnets produced by spark plasma sintering method. J. Alloys Compd. 2019, 773, 1018–1022. [Google Scholar] [CrossRef]
  84. Tozman, P.; Sepehri-Amin, H.; Zhang, L.T.; Wang, J.; Xu, X.D.; Hono, K. An alternative approach to the measurement of anisotropy field—Single grain extraction. J. Magn. Magn. Mater. 2020, 494, 165747-1–165747-4. [Google Scholar] [CrossRef]
  85. Dirba, I.; Harashima, Y.; Sepehri-Amin, H.; Ohkubo, T.; Miyake, T.; Hirosawa, S.; Hono, K. Thermal decomposition of ThMn12-type phase and its optimum stabilizing elements in SmFe12-based alloys. J. Alloys Compd. 2020, 813, 152224. [Google Scholar] [CrossRef]
  86. Hadjipanayis, G.C.; Gabay, A.M.; Schonhobel, A.M.; Martín-Cid, A.; Barandiaran, J.M.; Niarchos, D. ThMn12-Type Alloys for permanent magnets. Eng. J. 2020, 6, 141–147. [Google Scholar] [CrossRef]
  87. Kuno, T.; Suzuki, S.; Urushibata, K.; Kobayashi, K.; Sugimoto, S. Experimental determination of the saturated polarization and the anisotropy field in ThMn12-type magnets using the LAFS method. J. Magn. Magn. Mater. 2020, 498, 166114. [Google Scholar] [CrossRef]
  88. Tozman, P.; Sepehri-Amin, H.; Ohkubo, T.; Hono, K. Intrinsic magnetic properties of (Sm,Gd)Fe12-based compounds with minimized addition of Ti. J. Alloys Compd. 2021, 855, 157491. [Google Scholar] [CrossRef]
  89. Dirba, I.; Sepehri-Amin, H.; Choi, I.-J.; Choi, J.-H.; Uh, H.S.; Kim, T.-H.; Kwon, S.-J.; Ohkubo, T.; Hono, K. SmFe12-based hard magnetic alloys prepared by reduction-diffusion process. J. Alloys Compd. 2021, 861, 157993. [Google Scholar] [CrossRef]
  90. Makurenkova, A.; Ogawa, D.; Tozman, P.; Okamoto, S.; Nikitin, S.; Hirosawa, S.; Hono, K.; Takahashi, Y.K. Intrinsic hard magnetic properties of Sm(Fe,Co)12xTix compound with ThMn12 structure. J. Alloys Compd. 2021, 861, 158477. [Google Scholar] [CrossRef]
  91. Yang, Y.-C.; Hong, S.; Zhang, Z.; Tong, L.; Gao, J.-L. Crystallographic and magnetic properties of substituted YTi(Fe1xTx)11. Solid State Commun. 1988, 68, 175–179. [Google Scholar] [CrossRef]
  92. Harashima, Y.; Terakura, K.; Kino, H.; Ishibashi, S.; Miyake, T. First-principles study on stability and magnetism of NdFe11M and NdFe11MN for M = Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn. J. Appl. Phys. 2016, 120, 203904. [Google Scholar] [CrossRef] [Green Version]
  93. Sato, T.; Ohsuna, T.; Yano, M.; Kato, A.; Kaneko, Y. Permanent magnetic properties of NdFe12Nx sputtered films epitaxially grown on V buffer layer. J. Appl. Phys. 2017, 122, 053903. [Google Scholar] [CrossRef]
  94. De Boer, F.; Huang, Y.-K.; De Mooij, D.; Buschow, K. Magnetic properties of a series of novel ternary intermetallics (RFe10V2). J. Less Common. Met. 1987, 135, 199–204. [Google Scholar] [CrossRef]
  95. Hellholdt, R.B.; Vleggaar, J.J.M.; Buschow, K.H.K. Note of the crystallographic and magnetic structure of YFe10V2. J. Less Common. Met. 1988, 138, L11–L14. [Google Scholar] [CrossRef]
  96. Ding, J.; Rosenberg, M. Magnetic hardening of melt-spun and crystallized Sm-Fe-V and Sm-(Fe,Co)-V alloys. J. Less. Common. Met. 1990, 161, 369–374. [Google Scholar] [CrossRef]
  97. Jurczyk, M. Magnetic and crystallographic properties of SmFe10xCoxV2 compounds. J. Less. Common. Met. 1990, 162, 149–154. [Google Scholar] [CrossRef]
  98. Haije, W.G.; Spijkerman, J.; de Boer, F.R.; Bakker, K.; Buschow, K.H.J. Magnetic structure of rare earth compounds of the type RFe10V2. J. Less. Common. Met. 1990, 162, 285–295. [Google Scholar] [CrossRef]
  99. Schultz, L.; Schnitzke, K.; Wecker, J. High coercivity in mechanically alloyed Sm-Fe-V magnets with a ThMn12 crystal structure. Appl. Phys. Lett. 1990, 56, 868–870. [Google Scholar] [CrossRef]
  100. Zhong, H.P.; de Boer, F.R.; de Mooij, D.B.; Buschow, K.H.J. Magnetic coupling in the tetragonal rare earth iron compounds of the R(Fe,V)12. J. Less. Common. Met. 1990, 163, 123–132. [Google Scholar] [CrossRef]
  101. Yang, J.; Dong, S.-Z.; Yang, Y.-C.; Cheng, B.P. Structural and magnetic properties of RFe10.5V1.5Nx, Structural and magnetic properties of RFe10.5V1.5Nx. J. Appl. Phys. 1994, 75, 3013–3016. [Google Scholar]
  102. Sugimoto, S.; Shimono, T.; Nakamura, H.; Kagotani, T.; Okada, M.; Homma, M. Phase relation of the Sm-Fe-V alloys around compound Sm3(Fe,V)29. Mater. Trans. JIM 1996, 37, 494–498. [Google Scholar] [CrossRef] [Green Version]
  103. Schönhöbel, A.M.; Madugundo, R.; Gabay, A.M.; Barandiarán, J.M.; Hadjipanayis, C. Sm-Fe-V based 1:12 bulk magnets. J. Alloys Compd. 2019, 791, 1122–1127. [Google Scholar] [CrossRef] [Green Version]
  104. Schönhöbel, A.M.; Madugundo, R.; Vekilova, O.Y.; Eriksson, O.; Herper, H.C.; Barandiarán, J.M.; Hadjipanayis, G.C. Intrinsic magnetic properties of SmFe12−xVx alloys with reduced V-concentration. J. Alloys Compd. 2019, 786, 969–974. [Google Scholar] [CrossRef] [Green Version]
  105. Wang, X.Z.; Chevalier, B.; Berlureau, T.; Etourneau, J.; Coey, J.M.D.; Cadogan, J.M. Iron-rich pseudobinary alloys with the ThMn12 structure obtained by melt spinning Gd(FenAl12−n), n = 6, 8, 10. J. Less.-Common. Met. 1988, 138, 235–240. [Google Scholar] [CrossRef] [Green Version]
  106. Simon, D.; Wuest, H.; Koehler, T.; Senyshyn, A.; Ehrenberg, H.; Gutfleisch, O. The quaternary system Sm-Fe-Mo-Al and the effect of Al substitution on magnetic and structural properties of its ThMn12 phase. J. Alloys Compd. 2019, 770, 301–307. [Google Scholar] [CrossRef]
  107. Zhou, C.; Pinkerton, F.E.; Herbst, J.F. High Curie temperature of Co-Fe-Si compound with ThMn12 structure. Scr. Mater. 2015, 95, 66–69. [Google Scholar] [CrossRef]
  108. Gabay, A.M.; Cabassi, R.; Fabbrici, S.; Albertini, F.; Hadjipanayis, G.C. Structure and permanent magnet properties of Zr1−xRxFe10Si2 alloys with R = Y, La, Ce, Pr and Sm. J. Alloys Compd. 2016, 683, 271–275. [Google Scholar] [CrossRef] [Green Version]
  109. Gabay, A.M.; Hadjipanayis, G.C. ThMn12-type structure and uniaxial magnetic anisotropy in ZrFe10Si2 and Zr1−xCexFe10Si2 alloys. J. Alloys Compd. 2016, 657, 133–137. [Google Scholar] [CrossRef] [Green Version]
  110. Gjoka, M.; Sarafidis, C.; Psycharis, V.; Devlin, E.; Niarchos, D.; Hadjipanayis, G. Structure and magnetic properties of Sm1−xZrxFe10Si2 (x = 0.9–0.6) alloys. IOP Conf. Ser. J. Phys. Conf. Ser. 2017, 903, 12033. [Google Scholar]
  111. Gjoka, M.; Psycharis, V.; Devlin, E.; Niarchos, D.; Hadjipanayis, G. Effect of Zr substitution on the structural and magnetic properties of the series Nd1−xZrxFe10Si2 with the ThMn12 type structure. J. Alloys Compd. 2016, 687, 240–245. [Google Scholar] [CrossRef]
  112. Gabay, A.M.; Hadjipanayis, G.C. Mechanochemical synthesis of magnetically hard anisotropic RFe10Si2 powders with R representing combinations of Sm, Ce and Zr. J. Magn. Magn. Mater. 2017, 422, 43–48. [Google Scholar] [CrossRef] [Green Version]
  113. Barandiaran, J.M.; Martin-Cid, A.; Schönhöbel, A.M.; Garitaonandia, J.S.; Gjoka, M.; Niarchos, D.; Makridis, S.; Pasko, A.; Aubert, A.; Mazaleyrat, F.; et al. Nitrogenation and sintering of (Nd-Zr)Fe10Si2 tetragonal compounds for permanent magnets applications. J. Alloys Compd. 2019, 784, 996–1002. [Google Scholar] [CrossRef] [Green Version]
  114. Sakurada, S.; Tsutai, A.; Sahashi, M. A study on the formation of ThMn12 and NaZn13 structures in RFe10Si2. J. Alloys Compd. 1992, 187, 67–71. [Google Scholar] [CrossRef]
  115. Yücel, A.; Elmali, A.; Çakır, R.; Elerman, Y. Competing magnetic interactions in Dy1xSmxFe10Si2 compounds. Solid State Commun. 2005, 135, 745–748. [Google Scholar] [CrossRef]
  116. Muller, A. Magnetic material R,Fe,Mo,(Co) with ThMn12 structure. J. Appl. Phys. 1988, 64, 249–251. [Google Scholar] [CrossRef]
  117. Kou, X.C.; Grössinger, R.; Wiesinger, G.; Liu, J.P.; de Boer, F.R.; Kleinschroth, I.; Kronmüller, H. Intrinsic magnetic properties of RFe10Mo2 compounds (R = Y, Pr, Nd, Sm, Gd, Tb, Dy, Ho, Er, or Tm). Phys. Rev. B 1995, 51, 8254–8265. [Google Scholar] [CrossRef] [Green Version]
  118. Loong, C.-K.; Short, S.M.; Lin, J.; Ding, Y. A study of nitrogenation of a NdFe12−xMox compound by in situ neutron powder diffraction. J. Appl. Phys. 1998, 83, 6926–6928. [Google Scholar] [CrossRef]
  119. Zinkevitch, M.; Mattern, N.; Wetzig, K. Structural properties of the ThMn12-type nonstoichiometric solid solutions in the Fe-Gd-Mo system. J. Alloys Compd. 1998, 268, 155–160. [Google Scholar] [CrossRef]
  120. Du, H.L.; Han, J.Z.; Zhang, W.Y.; Wang, C.S.; Wang, W.C.; Liu, S.Q.; Chen, H.; Zhang, X.D.; Yang, Y.C. Determination of the zirconium site in zirconium-substituted Nd(Fe, Mo, Zr)12 compounds. J. Magn. Magn. Mater. 2004, 283, 316–321. [Google Scholar] [CrossRef]
  121. Fuquan, B.; Wang, J.L.; Tegus, O.; Dagula, W.; Tang, N.; Yang, F.M.; Wu, G.H.; Brück, E.; de Boer, F.R.; Buschow, K.H.J. Phase formation and magnetic properties of YFe12xNbx (x = 0.70–0.90) compounds. J. Magn. Magn. Mater. 2005, 290–291 Pt 2, 1192–1194. [Google Scholar] [CrossRef]
  122. Sakurada, S.; Tsutai, A.; Hirai, T.; Yanagida, Y.; Sahashi, M.; Abe, S.; Kaneko, T. Structural and magnetic properties of rapidly quenched (R,Zr)(Fe,Co)10Nx (R = Nd, Sm). J. App. Phys. 1996, 79, 4611–4613. [Google Scholar] [CrossRef]
  123. Sakuma, N.; Suzuki, S.; Kuno, T.; Urushibata, K.; Kobayashi, K.; Yano, M.; Kato, A.; Manabe, A. Influence of Zr substitution on the stabilization of ThMn12-type (Nd1-αZrα)(Fe0.75Co0.25)11.25Ti0.75N1.2−1.4 (α = 0–0.3) compounds. AIP Adv. 2016, 6, 56023. [Google Scholar] [CrossRef] [Green Version]
  124. Hirayama, Y.; Takahashi, Y.K.; Hirosawa, S.; Hono, K. NdFe12Nx hard-magnetic compound with high magnetization and anisotropy field. Scr. Mater. 2015, 95, 70–72. [Google Scholar] [CrossRef] [Green Version]
  125. Coehoorn, R. Electronic structure and magnetism of transition-metal-stabilized YFe12xMx intermetallic compounds. Phys. Rev. B 1990, 41, 11790–11797. [Google Scholar] [CrossRef]
  126. Hirayama, Y.; Takahashi, Y.; Hirosawa, S.; Hono, K. Intrinsic hard magnetic properties of Sm(Fe1−xCox)12 compound with the ThMn12 structure. Scr. Mater. 2017, 138, 62–65. [Google Scholar] [CrossRef]
  127. Sun, H.; Tomida, T.; Hirosawa, S. Magnetic properties and microstructure studies of Sm-Fe magnetic thin films. J. Appl. Phys. 1997, 81, 328–334. [Google Scholar] [CrossRef]
  128. Ogawa, D.; Xu, X.D.; Takahashi, Y.K.; Ohkubo, T.; Hirosawa, S.; Hono, K. Emergence of coercivity in Sm(Fe0.8Co0.2)12 thin films via eutectic alloy grain boundary infiltration. Scr. Mater. 2019, 164, 140–144. [Google Scholar] [CrossRef]
  129. Ogawa, D.; Yoshioka, T.; Xu, X.D.; Takahashi, Y.K.; Tsuchiura, H.; Ohkubo, T.; Hirosawa, S.; Hono, K. Magnetic anisotropy constants of ThMn12-type Sm (Fe1−xCox)12 compounds and their temperature dependence. J. Magn. Magn. Mater. 2020, 497, 165965. [Google Scholar] [CrossRef]
  130. Hono, K.; Sepehri-Amin, H. Prospect for HRE-free high coercivity Nd-Fe-B permanent magnets. Scr. Mater. 2018, 151, 6–13. [Google Scholar] [CrossRef]
  131. Liang, J.K.; Huang, Q.; Santoro, A.; Wang, L.; Yang, F.M. Magnetic structure and site occupancies in YFe11xCoxTi (x = 1, 3, 7, 9). J. Appl. Phys. 1999, 86, 2155–2160. [Google Scholar] [CrossRef]
  132. Chen, N.-X.; Hao, S.-Q.; Wu, Y.; Shen, J. Phase stability and site preference of Sm(Fe,T)12. J. Magn. Magn Mater. 2001, 233, 169–180. [Google Scholar]
  133. Ke, L.; Johnson, D.D. Intrinsic magnetic properties in R(Fe1−xCox)11TiZ (R = Y and Ce; Z = H, C, and N). Phys. Rev. B 2016, 94, 244231. [Google Scholar] [CrossRef] [Green Version]
  134. Körner, W.; Krugel, G.; Elsässer, C. Theoretical screening of intermetallic ThMn12-type phases for new hard-magnetic compounds with low rare earth content. Sci. Rep. 2016, 6, 24686. [Google Scholar] [CrossRef] [PubMed]
  135. Butcher, T.A.; Körner, W.; Krugel, G.; Elsässer, C. Dependence of magnetization and magnetocrystalline anisotropy on site distribution of alloying elements in RE-TM phases with ThMn12 structure. J. Magn. Magn. Mater. 2017, 441, 1–5. [Google Scholar] [CrossRef]
  136. Körner, W.; Krugel, G.; Urban, D.F.; Elsässer, C. Screening of rare-earth-lean intermetallic 1-11 and 1-11-X compounds of YNi9In2-type for hard-magnetic applications. Scr. Mater. 2018, 154, 295–299. [Google Scholar] [CrossRef]
  137. Skelland, C.; Ostler, T.; Westmoreland, S.C.; Evans, R.F.L.; Chantrell, R.W.; Yano, M.; Shoji, T.; Manabe, A.; Kato, A.; Ito, M.; et al. Probability distribution of dubstituted titanium in RT12 (R = Nd and Sm; T = Fe and Co) structures. IEEE Trans. Magn. 2018, 54, 2103405. [Google Scholar] [CrossRef]
  138. Miyake, T.; Terakura, K.; Harashima, Y.; Kino, H.; Ishibashi, S. First-Principles Study of Magnetocrystalline Anisotropy and Magnetization in NdFe12, NdFe11Ti, and NdFe11TiN. J. Phys. Soc. Jpn. 2014, 83, 43702. [Google Scholar] [CrossRef]
  139. Odkhuu, D.; Ochirkhuyag, T.; Hong, S.C. Enhancing energy product and thermal stability of SmFe12 by interstitial doping. Phys Rev. Appl. 2020, 13, 54076. [Google Scholar] [CrossRef]
  140. Harashima, Y.; Terakura, K.; Kino, H.; Ishibashi, S.; Miyake, T. Nitrogen as the best interstitial dopant among X = B, C, N, O, and F for strong permanent magnet NdFe11TiX: First-principles study. Phys. Rev. B 2015, 92, 184426. [Google Scholar] [CrossRef] [Green Version]
  141. Harashima, Y.; Terakura, K.; Kino, H.; Ishibashi, S.; Miyake, T. First-Principles Study of Structural and Magnetic Properties of R(Fe,Ti)12 and R(Fe,Ti)12N (R = Nd, Sm, Y). JPS Conf. Proc. 2015, 5, 11021. [Google Scholar]
  142. Matsumoto, M.; Akai, H.; Harashima, Y.; Doi, S.; Miyake, T. Relevance of 4f-3d exchange to finite-temperature magnetism of rare-earth permanent magnets: An ab-initio-based spin model approach for NdFe12N. J. Appl. Phys. 2016, 119, 213901. [Google Scholar] [CrossRef] [Green Version]
  143. Fukazawa, T.; Akai, H.; Harashima, Y.; Miyake, T. First-principles study of intersite magnetic couplings in NdFe12 and NdFe12X (X = B, C, N, O, F). J. Appl. Phys. 2017, 122, 53901. [Google Scholar] [CrossRef] [Green Version]
  144. Fukazawa, T.; Akai, H.; Harashima, Y.; Miyake, T. First-principles study of intersite magnetic couplings and Curie temperature in RFe12−xCrx (R = Y, Nd, Sm). J. Phys. Soc. Jpn. 2018, 87, 44706. [Google Scholar] [CrossRef] [Green Version]
  145. Harashima, Y.; Fukazawa, T.; Kino, H.; Miyake, T. Effect of R-site substitution and pressure on stability of RFe12 a first-principles study. J. Appl. Phys. 2018, 124, 163902. [Google Scholar] [CrossRef] [Green Version]
  146. Fukazawa, T.; Harashima, Y.; Hou, Z.; Miyake, T. Bayesian optimization of chemical composition: A comprehensive framework and its application to RFe12-type magnet compounds. Phys. Rev. Mater. 2019, 3, 53807. [Google Scholar] [CrossRef] [Green Version]
  147. Matsumoto, M.; Hawai, T.; Ono, K. (Sm, Zr)Fe12−xMx (M = Zr, Ti, Co) for permanent-magnet applications: Ab initio material design integrated with experimental characterization. Phys. Rev. Appl. 2020, 13, 64028. [Google Scholar] [CrossRef]
  148. Dasmahapatra, A.; Martinez-Casado, R.; Romero-Muñiz, C.; Sgroi, M.F.; Ferrari, A.M.; Maschio, L. Doping the permanent magnet CeFe11Ti with Co and Ni using ab-initio density functional methods. Phys. B 2021, 620, 413241. [Google Scholar] [CrossRef]
  149. Söderlind, P.; Landa, A.; Locht, I.L.M.; Åberg, D.; Kvashnin, Y.; Pereiro, M.; Däne, M.; Turchi, P.E.A.; Antropov, V.P.; Eriksson, O. Prediction of the new efficient permanent magnet SmCoNiFe3. Phys. Rev. B. 2017, 96, 100404. [Google Scholar] [CrossRef] [Green Version]
  150. Landa, A.; Söderlind, P.; Parker, D.; Åberg, D.; Lordi, V.; Perron, A.; Turchi, P.E.A.; Chouhan, R.K.; Paudyal, D.; Lograsso, T.A. Thermodynamics of the SmCo5 compound doped with Fe and Ni: An ab initio study. J. Alloys Compd. 2018, 765, 659–663. [Google Scholar] [CrossRef] [Green Version]
  151. Landa, A.; Söderlind, P.; Moore, E.E.; Perron, A. Thermodynamics and magnetism of YCo5 compound doped with Fe and Ni: An ab initio study. Appl. Sci. 2020, 10, 6037. [Google Scholar] [CrossRef]
  152. Suski, W.; Wochowski, K.; Mydlarz, T.; Belan, B. Magnetic properties of SmMn12−xNix system. Phys. Stat. Sol. 2003, 196, 290–293. [Google Scholar] [CrossRef]
  153. Suski, W.; Belan, B.; Mydlarz, T.; Wochowski, K. Magnetic properties of NdMn12−xNix alloys. J. Alloys Compd. 2004, 367, 215–218. [Google Scholar] [CrossRef]
  154. Söderlind, P.; Eriksson, O.; Johansson, B.; Albers, R.C.; Boring, A.M. Spin and orbital magnetism in Fe-Co and Co-Ni alloys. Phys. Rev. B 1992, 66, 12911–12916. [Google Scholar] [CrossRef] [PubMed]
  155. Antonov, V.N.; Harmon, B.N.; Yaresko, A.N. Electronic structure of mixed-valence semiconductors in the LSDA+U approximation. I. Sm monochalcogenides. Phys. Rev. B 2002, 45, 165208. [Google Scholar]
  156. Grånäs, O.; Di Marco, I.; Thunström, P.; Nordström, L.; Eriksson, O.; Björkman, T.; Wills, J.M. Charge self-consistent dynamical mean-field theory based on the full-potential linear muffin-tin orbital method: Methodology and applications. Comput. Mater. Sci. 2012, 55, 295–302. [Google Scholar] [CrossRef] [Green Version]
  157. Locht, I.L.M.; Kvashnin, Y.O.; Rodrigues, D.C.M.; Pereiro, M.; Bergman, A.; Bergqvist, L.; Lichtenstein, A.I.; Katsnelson, M.I.; Delin, A.; Klautau, A.; et al. Standard model of the rare earths analyzed from the Hubbard I approximation. Phys. Rev. B 2016, 94, 85137. [Google Scholar] [CrossRef] [Green Version]
  158. Koehler, W.C.; Moon, R.M. Magnetic Structures of Samarium. Phys. Rev. Lett. 1972, 29, 1468–1472. [Google Scholar] [CrossRef]
  159. Vitos, L. Total-energy method based on the exact muffin-tin orbital theory. Phys. Rev. B 2001, 64, 14107. [Google Scholar] [CrossRef]
  160. Vitos, L. Computational Quantum Mechanics for Materials Engineers: The EMTO Method and Application; Springer: London, UK, 2007. [Google Scholar]
  161. Kollar, J.; Vitos, L.; Skriver, H.L. From ASA towards the Full Potential. In Lecture Notes in Physics; Dreyssé, H., Ed.; Springer: Berlin/Heidelberg, Germany, 2000; pp. 85–113. [Google Scholar]
  162. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  163. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  164. Pourovskii, L.V.; Ruban, A.V.; Vitos, L.; Ebert, H.; Johansson, B.; Abrikosov, I.A. Fully relativistic spin-polarized exact muffin-tin-orbital method. Phys. Rev. B 2005, 71, 94415. [Google Scholar] [CrossRef]
  165. Faulkner, J.S. The modern theory of alloys. Prog. Mater. Sci. 1982, 27, 1–187. [Google Scholar] [CrossRef]
  166. Vitos, L.; Abrikosov, I.A.; Johansson, B. Anisotropic lattice distortions in random alloys from first-principles theory. Phys. Rev. Lett. 2001, 87, 156401. [Google Scholar] [CrossRef] [PubMed]
  167. Staunton, J.; Gyorffy, B.; Pindor, A.; Stocks, G.M.; Winter, H. The “disordered local moment” picture of itinerant magnetism at finite temperatures. J. Magn. Magn. Mater. 1984, 45, 15–22. [Google Scholar] [CrossRef]
  168. Gyorffy, B.L.; Pindor, A.; Staunton, J.; Stocks, G.M.; Winter, H. A first-principles theory of ferromagnetic phase transitions in metals. J. Phys. F Met. Phys. 1985, 15, 1337–1386. [Google Scholar] [CrossRef]
  169. Murnaghan, F.D. The compressibility of media under extreme pressures. Proc. Natl. Acad. Sci. USA 1944, 30, 244–247. [Google Scholar] [CrossRef] [Green Version]
  170. Andersen, O.K. Linear methods in band theory. Phys. Rev. B 1975, 12, 3060–3083. [Google Scholar] [CrossRef] [Green Version]
  171. Söderlind, P.; Turchi, P.E.A.; Landa, A.; Lordi, V. Ground-state properties of rare-earth metals: An evaluation of density-functional theory. J. Phys. Condens. Matter 2014, 26, 416001. [Google Scholar] [CrossRef]
  172. Söderlind, P. Delocalization and phase transitions in Pr: Theory. Phys. Rev. B 2002, 65, 115105. [Google Scholar] [CrossRef]
  173. Wills, J.M.; Alouani, M.; Andersson, P.; Delin, A.; Eriksson, O.; Grechnyev, O. Full-Potential Electronic Structure Method; Springer Series in Solid-State Science; Springer: Berlin/Heidelberg, Germany, 2010; Volume 167. [Google Scholar]
  174. von Barth, U.; Hedin, L. A local exchange-correlation potential for the spin polarized case. J. Phys. C 1972, 5, 1629–1642. [Google Scholar] [CrossRef]
  175. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  176. Wang, D.; Liou, S.H.; He, P.; Sellmyer, D.J.; Hadjipanayis, G.C.; Zhang, Y. SmFe12 and SmFe12Nx films fabricated by sputtering. J. Magn. Magn. Mater. 1993, 124, 62–68. [Google Scholar] [CrossRef]
  177. Harashima, Y.; Fukazawa, T.; Miyake, T. Cerium as a possible stabilizer of ThMn12-type iron-based compounds: A first-principles study. Scr. Mater. 2020, 179, 12–15. [Google Scholar] [CrossRef] [Green Version]
  178. Sato, K.; Katayama-Yoshida, H.; Dederichs, P. Dilute magnetic semiconductors. Newsletter 2005, 70, 93–110. [Google Scholar]
  179. Ma, D.; Grabowski, B.; Körmann, F.; Neugebauer, J.; Raabe, D. Ab initio thermodynamics of the CoCrFeMnNi high entropy alloy: Importance of entropy contributions beyond the configurational one. Acta Mater. 2015, 100, 90–97. [Google Scholar] [CrossRef]
  180. Skomski, R.; Coey. J.M.D. Magnetic anisotropy—How much is enough for a permanent magnet? Scr. Mater. 2016, 112, 3–8. [Google Scholar] [CrossRef] [Green Version]
  181. Miyake, T.; Harashima, Y.; Fukazawa, T.; Akai, H. Understanding and optimization of hard magnetic compounds from first principles. Sci. Technol. Adv. Mater. 2021, 22, 543–556. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of the SmFe12 (ThMn12-type) compound sketched using Vesta software [175]. The larger pink spheres denote the Sm atoms at Wyckoff position 2a, while the smaller white, gray, and black spheres are the Fe atoms at Wyckoff positions 8f, 8i, and 8j, respectively. Adapted with permission from Ref. [139]. Reuse and Permissions License Number: RNP/22/MAY/053350.
Figure 1. Crystal structure of the SmFe12 (ThMn12-type) compound sketched using Vesta software [175]. The larger pink spheres denote the Sm atoms at Wyckoff position 2a, while the smaller white, gray, and black spheres are the Fe atoms at Wyckoff positions 8f, 8i, and 8j, respectively. Adapted with permission from Ref. [139]. Reuse and Permissions License Number: RNP/22/MAY/053350.
Applsci 12 04860 g001
Figure 2. The formation energy of Sm(Fe1−xCox)12 alloys.
Figure 2. The formation energy of Sm(Fe1−xCox)12 alloys.
Applsci 12 04860 g002
Figure 3. The formation energy values of the SmNi4(Fe1−xCox)8 alloys, where 4 Ni atoms occupy the 8j sublattice and 8 (Fe1−xCox) atoms are randomly distributed on the 8i and 8f sublattices.
Figure 3. The formation energy values of the SmNi4(Fe1−xCox)8 alloys, where 4 Ni atoms occupy the 8j sublattice and 8 (Fe1−xCox) atoms are randomly distributed on the 8i and 8f sublattices.
Applsci 12 04860 g003
Figure 4. The saturation magnetization of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Figure 4. The saturation magnetization of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Applsci 12 04860 g004
Figure 5. The maximum energy product of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Figure 5. The maximum energy product of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Applsci 12 04860 g005
Figure 6. The Curie temperature of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Figure 6. The Curie temperature of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Applsci 12 04860 g006
Figure 7. The samarium magnetic moments for various theoretical models (see main text) of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Figure 7. The samarium magnetic moments for various theoretical models (see main text) of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Applsci 12 04860 g007
Figure 8. The anisotropy field of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Figure 8. The anisotropy field of the SmNi4(Fe1−xCox)8 alloys, x = 0.0, 0.1, 0.2, 1.0.
Applsci 12 04860 g008
Figure 9. Calculated (black circles) formation energies of the SmFe12, SmCo12, and SmNi12 compounds.
Figure 9. Calculated (black circles) formation energies of the SmFe12, SmCo12, and SmNi12 compounds.
Applsci 12 04860 g009
Table 1. Saturation magnetization, anisotropy field, Curie temperature, and maximum energy product values of 1:12 magnets.
Table 1. Saturation magnetization, anisotropy field, Curie temperature, and maximum energy product values of 1:12 magnets.
Material μ 0 M s T μ 0 H a T T c K   B H m a x k J m 3 References
Nd2Fe14B11.617.6588515[7,8,9,10,11,12,13,14]
SmFe11Ti11.14–1.228.3–10.5584258[23,34]
(Sm0.8Zr0.2)(Fe0.75Co0.25)11.5Ti0.51.587.4880495[69]
(Sm0.8Y0.2)(Fe0.80Co0.20)11.5Ti0.51.5011.0820447[73,78]
Sm0.94(Fe0.81Co0.19)11Ti1.081.4310.9800406[75]
(Sm0.77Zr0.24)(Fe0.80Co0.19)11.5Ti0.651,538.4830465[75]
(Sm0.92Zr0.08)(Fe0.75Co0.25)11.35Ti0.651.47>9.0843429[76]
Sm(Fe0.8Co0.2)121.7812.0859630[126]
(Sm0.74Zr0.26)(Fe0.8Co0.2)121.869.8875688[80,84]
Table 2. The decomposition energy of the reaction RE(Fe,Co,Ni)12RE2(Fe, Co)17 + TMs.
Table 2. The decomposition energy of the reaction RE(Fe,Co,Ni)12RE2(Fe, Co)17 + TMs.
MaterialE (mRy/atom)
SmFe12+7.79
SmCo12+2.13
SmNi4Fe8−1.29
SmNi4Co8−0.59
Table 3. Site-projected spin (m(s)) and orbital (m(o)) magnetic moments for the SmNi4Fe8, SmNi4(Fe.9Co.1)8, SmNi4(Fe.8Co.2)8, and SmNi4Co8 compounds, where 4 Ni atoms occupy the 8j sublattice and 8 (Fe,Co) atoms are equally distributed on the 8i and 8f sublattices: mtotal= 19.9349, 19.9778, 19.5601, and 13.5833 μB/f.u, respectively.
Table 3. Site-projected spin (m(s)) and orbital (m(o)) magnetic moments for the SmNi4Fe8, SmNi4(Fe.9Co.1)8, SmNi4(Fe.8Co.2)8, and SmNi4Co8 compounds, where 4 Ni atoms occupy the 8j sublattice and 8 (Fe,Co) atoms are equally distributed on the 8i and 8f sublattices: mtotal= 19.9349, 19.9778, 19.5601, and 13.5833 μB/f.u, respectively.
Sm1(2a)Fe1(8f)/Co1(8f)Fe2(8i)/Co2(8i)Ni3 (8j)
SmNi4Fe8
m(s) (μB)+3.7182−2.1482/-−2.1763/-−0.6144
m(o) (μB)−3.1649−0.0652/-−0.0658/-−0.0522
SmNi4(Fe.9Co.1)8
m(s) (μB)+3.6956−2.2330/−1.4267−2.2568/−1.3874−0.5977
m(o) (μB)−3.2281−0.0678/−0.0861−0.0693/−0.0800−0.0514
SmNi4(Fe.8Co.2)8
m(s) (μB)+3.7452−2.2866/−1.4492−2.2898/−1.3992−0.5936
m(o) (μB)−3.2376−0.0686/−0.0869−0.0698/−0.0809−0.0481
SmNi4Co8
m(s) (μB)+4.1589-/−1.5299-/−1.4692−0.5420
m(o) (μB)−2.8154-/−0.0899-/−0.0678−0.0329
Table 4. Atomic volume (Ω0), density (ρ), total moment (mtotal), saturation magnetization (Ms and μ0Ms), and maximum energy product (|BH|max) as calculated by the EMTO method and using the FPLMTO method with different approaches: 4f-band-scalar, 4f-band-SO, 4f-band-SO + OP, and SRM + OP (see main text for explanation of these models).
Table 4. Atomic volume (Ω0), density (ρ), total moment (mtotal), saturation magnetization (Ms and μ0Ms), and maximum energy product (|BH|max) as calculated by the EMTO method and using the FPLMTO method with different approaches: 4f-band-scalar, 4f-band-SO, 4f-band-SO + OP, and SRM + OP (see main text for explanation of these models).
MaterialTheoryΩ0
3)
Ρ 
( g c m 3 )
m t o t a l  
μ B f . u .
M s  
M A m
μ 0 M s  
T
B H m a x  
k J m 3
SmFe12EMTO13.197.6925.71.391.75608.0
SmCo12EMTO12.428.8217.41.001.26315.2
SmNi4Fe8EMTO12.988.1919.91.101.38377.6
4f-band-scalar12.848.2816.20.901.14254.4
4f-band-SO12.848.2818.71.041.31339.4
4f-band-SO + OP12.908.2420.11.121.41388.6
SRM + OP12.988.1922.51.241.57481.3
SmNi4(Fe0.9Co0.1)8EMTO12.928.2519.91.101.39382.3
4f-band-scalar12.798.3315.60.871.09237.5
4f-band-SO12.798.3318.11.011.27319.7
4f-band-SO + OP12.858.2919.61.081.37371.2
SRM + OP12.938.2422.01.211.53462.5
SmNi4(Fe0.8Co0.2)8EMTO12.898.3019.61.081.36368.2
4f-band-scalar12.748.4015.00.841.06221.7
4f-band-SO12.748.4017.40.971.23298.2
4f-band-SO + OP12.808.3619.01.061.33352.6
SRM + OP12.878.3121.41.181.42441.7
SmNi4Co8EMTO12.478.7713.60.780.98189.7
4f-band-scalar12.338.879.600.560.7097.08
4f-band-SO12.338.8711.80.690.86146.3
4f-band-SO + OP12.378.8413.60.790.99193.1
SRM + OP12.498.7616.10.921.16265.6
Table 5. Calculated atomic volume, magnetic anisotropy energy, first anisotropy constant, anisotropy field, and magnetic hardness parameter for the SmNi4(Fe1−xCox)8 alloys, where x = 0, 0.1, 0.2, and 1.
Table 5. Calculated atomic volume, magnetic anisotropy energy, first anisotropy constant, anisotropy field, and magnetic hardness parameter for the SmNi4(Fe1−xCox)8 alloys, where x = 0, 0.1, 0.2, and 1.
CompoundΩ0 M A E   m e V f . u . K 1 M J m 3   μ0Ha (T)κ
SmNi4Fe812.983.983.786.091.40
SmNi4(Fe0.9Co0.1)812.935.094.858.021.62
SmNi4(Fe0.8Co0.2)812.876.496.2210.541.88
SmNi4Co812.499.539.4120.132.98
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Landa, A.; Söderlind, P.; Moore, E.E.; Perron, A. Thermodynamics and Magnetism of SmFe12 Compound Doped with Co and Ni: An Ab Initio Study. Appl. Sci. 2022, 12, 4860. https://doi.org/10.3390/app12104860

AMA Style

Landa A, Söderlind P, Moore EE, Perron A. Thermodynamics and Magnetism of SmFe12 Compound Doped with Co and Ni: An Ab Initio Study. Applied Sciences. 2022; 12(10):4860. https://doi.org/10.3390/app12104860

Chicago/Turabian Style

Landa, Alexander, Per Söderlind, Emily E. Moore, and Aurélien Perron. 2022. "Thermodynamics and Magnetism of SmFe12 Compound Doped with Co and Ni: An Ab Initio Study" Applied Sciences 12, no. 10: 4860. https://doi.org/10.3390/app12104860

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