thanks: [email protected]

Fast ab initio design of high-entropy magnetic thin films

Dinesh Bista Department of Physics, Georgetown University, Washington, D.C. 20057, USA    Willie B. Beeson Department of Physics, Georgetown University, Washington, D.C. 20057, USA    Turbasu Sengupta Department of Physics, Virginia Commonwealth University, Richmond, VA, 23284, USA    Jerome Jackson Scientific Computing Department, STFC Daresbury Laboratory, Warrington WA4 4AD, United Kingdom    Shiv N Khanna Department of Physics, Virginia Commonwealth University, Richmond, VA, 23284, USA    Kai Liu Department of Physics, Georgetown University, Washington, D.C. 20057, USA    Gen Yin Department of Physics, Georgetown University, Washington, D.C. 20057, USA
Abstract

We show that the magnetic properties of high-entropy alloys (HEAs) can be captured by ab initio calculations within the coherent potential approximation, where the atomic details of the high-entropy mixing are considered as an effective medium that possesses the translational symmetry of the lattice. This is demonstrated using the face-centered cubic (FCC) phase of FeCoNiMnCu and the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase of (FeCoNiMnCu)Pt by comparing the density functional theory (DFT) results with the experimental values. Working within the first Brillouin zone and the primitive unit cell, we show that DFT can capture the smooth profile of magnetic properties such as the saturation magnetization, the Curie temperature and the magnetic anisotropy, using only a sparse set of sampling points in the vast compositional space. The smooth profiles given by DFT indeed follow the experimental trend, demonstrating the promising potential of using machine learning to explore the magnetic properties of HEAs, by establishing reasonably large datasets with high-throughput calculations using density-functional theory.

High-entropy alloys (HEA) are a category of solids with a long-range ordered crystal structure, whereas the atomic sites are randomly occupied by the homogeneous, random mixing of five or more different elements[1, 2, 3, 4, 5, 6, 7, 8, 9, 10]. After this mixing, the change in the Gibbs free energy ΔG=ΔHTΔSΔ𝐺Δ𝐻𝑇Δ𝑆\Delta G=\Delta H-T\Delta Sroman_Δ italic_G = roman_Δ italic_H - italic_T roman_Δ italic_S is allowed to be negative even with a significant increase in the mixing enthalpy ΔH>0Δ𝐻0\Delta H>0roman_Δ italic_H > 0. This is enabled by the large configurational entropy gain ΔS>0Δ𝑆0\Delta S>0roman_Δ italic_S > 0. Such mixing can thus spontaneously stabilize a crystal structure that hosts many chemical bonds normally not favored in regular intermetallic alloys or ordered crystals[2]. Since the mixing is homogeneous and random, the stabilized crystal is often described as a solid solution[5]. Unlike regular ordered alloys, solid solutions host random local strains and lattice distortions because of the complex, multi-element composition. Therefore, atomic diffusion, dislocation, and grain boundary movements are impeded by numerous obstacles. This leads to increased hardness, thermal stability, and reduced brittleness, while surprisingly maintaining good ductility[11, 12]. Similar to the intriguing mechanical properties, solid-solution phases of HEA are also known to possess high chemical stability[13, 14, 8]. Due to the uniform mixing of the high-entropy atomic species, phase segregation in HEAs is more suppressed compared to regular alloys. This impedes the progress of electrochemical processes and thereby enhances the robustness against bimetallic corrosion. Due to these intriguing properties, HEAs have undergone extensive investigation, primarily focused on their enhanced mechanical and chemical properties [15, 16, 17, 6]. Recent developments in machine learning and autonomous workflow have brought both the experimental and theoretical research to a new level[18, 19, 20, 21].

In addition to mechanical and chemical robustness, HEAs can also exhibit exceptional magnetic properties[22, 23, 15, 24]. The ordered structures of HEAs naturally break the SO(3)𝑆𝑂3SO(3)italic_S italic_O ( 3 ) symmetry, allowing for good magnetic anisotropy. Also, the random local distortion of the lattice can serve as pinning sites, resulting in good permanent magnets with large coercivity[25]. Due to the compositional complexity, it is possible to modulate the key magnetic properties such as the saturation magnetization (MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT), the magnetic anisotropy energy density (MAE) and the Curie temperature (TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT) in a reasonably large range[26, 17]. This provides the opportunity to identify ideal media for advanced technologies such as heat- or microwave-assisted magnetic recording (HAMR and MAMR), where fine-tuning of MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT, MAE and TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT are needed to identify the best balance among the recording trilemma for each generation of the technology[27, 28, 29, 30, 31, 32, 33, 29]. This calls for the investigation of magnetic properties of thin-film HEAs in their vast, high-dimensional compositional space.

Here, we demonstrate the design principle for magnetic properties of thin-film HEAs using density-functional theory (DFT). By comparing to experimental results, we demonstrate that bulk DFT calculations using thin-film lattice constants can capture the smooth profile of magnetic properties in the high-entropy compositional space. Specifically, we demonstrate this design principle for the MAE of two thin-film HEAs that have been recently achieved experimentally: the face-centered cubic (FCC) phase of FeCoNiMnCu and the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase of (FeCoNiMnCu)Pt, where robust solid solutions with large MAE and exceptional long-range crystallization have been identified[34]. Particularly, we show that the first-principles calculation can be carried out efficiently at the level of coherent-potential approximation (CPA), a Green’s function-based theory for the electronic structure of a lattice featuring compositional and magnetic disorder, allowing for fast and accurate evaluation of magnetic properties of random alloys without the necessity of sampling different configurations using supercells[35, 36, 37]. This illustrates the possibility of using machine-learning models to make reasonable predictions once a large-scale database is established.

Refer to caption
Fig 1: (a) The Spectral function of L10-phase𝐿subscript10-phaseL1_{0}\thinspace\textrm{-phase}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT -phase FePt obtained by the Green’s function implementation of DFT in Questaal. Coherent potential approximation was used assuming no random mixing to resolve the band structure. An overall broadening energy of 0.001Ryd0.001Ryd0.001\thinspace\textrm{Ryd}0.001 Ryd was applied to visualize the spectral function. (b) The spectral function in the case of [Fe0.8CoNiMnCu0.2]Ptdelimited-[]subscriptFe0.8subscriptCoNiMnCu0.2Pt[\textrm{Fe}_{0.8}\textrm{CoNiMnCu}_{0.2}]\textrm{Pt}[ Fe start_POSTSUBSCRIPT 0.8 end_POSTSUBSCRIPT CoNiMnCu start_POSTSUBSCRIPT 0.2 end_POSTSUBSCRIPT ] Pt with the random mixing slightly turned on for Fe sites. (c) The VASP result of pure FePt with eigenstates projected to Fe and Pt orbitals. The weight of the projection is denoted by the color scale. The energy cutoff was set to 400eV400eV400\thinspace\textrm{eV}400 eV, and the DFT-D3 dispersion correction with Becke-Johnson dam** was used[38, 39, 40]. A ΓΓ\Gammaroman_Γ-centered 12×12×91212912\times 12\times 912 × 12 × 9 mesh was used for the k-space integration. The convergence criterion was set to 107eV.superscript107eV10^{-7}\thinspace\textrm{eV}.10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT eV .

The ab initio calculations in this work are performed using the Green’s function formulation of DFT and Linear Muffin Tin Orbitals within the atomic-sphere approximation (ASA). To model the irregularities in HEAs, we perform spin-polarized self-consistent calculations using the CPA implemented by the ‘lmgf’ package in the Questaal suite [41]. This approximation considers the irregularities as an averaged effective medium with the same translational symmetry of the lattice, allowing us to work within the first Brillouin zone. The electron density is obtained by integrating over a ΓΓ\Gammaroman_Γ-centered 16×16×1616161616\times 16\times 1616 × 16 × 16 k-mesh[42] and 25252525 energy points uniformly distributed along an elliptical contour of 0.40.40.40.4 eccentricity, covering the energy range of 0.85Ryd0.85Ryd0.85\thinspace\textrm{Ryd}0.85 Ryd below the Fermi level ϵFsubscriptitalic-ϵ𝐹\epsilon_{F}italic_ϵ start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. The value of ϵFsubscriptitalic-ϵ𝐹\epsilon_{F}italic_ϵ start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is determined by the charge-neutrality condition as implemented in the package. We determine the ground-state magnetic order by exploring all possible spin configurations, treating positive and negative spins along the quantization axis as two independent CPA species. The third-order potential function approximation (p3subscript𝑝3p_{3}italic_p start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) and the convergence criterion of 106Rydsuperscript106Ryd10^{-6}\thinspace\textrm{Ryd}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT Ryd are used. For the calculations of saturation magnetization, the magnetic moment per unit cell in the converged ground state is taken, and then converted to the units of emu/cm3superscriptemu/cm3\textrm{emu/cm}^{3}emu/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT or emu/g. Spin-orbit coupling is considered as a LSdelimited-⟨⟩𝐿𝑆\langle L\cdot S\rangle⟨ italic_L ⋅ italic_S ⟩ term added to the one-body Hamiltonian when calculating the MAE. The values of MAE are extracted by comparing the converged ground-state energy and another single-step energy using the converged moments in the previous step, but rotating the spin quantization axis from the easy axis to a perpendicular direction. The specific choices of quantization directions are made according to the setup of experiments.

We first calibrate our calculations by aligning Questaal results with other DFT implementations and experimental data. This comparison focuses on the well-known high anisotropy single crystal FePt in the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase. The Questaal calculation was carried out within CPA, assuming the 100%percent100100\%100 % occupation on both the Fe and Pt sites. The primitive cell of the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase is tetragonal, with Fe atoms occupying the corners and the Pt atom at the body center. We used a=2.69Å𝑎2.69Åa=2.69\thinspace\textrm{\AA}italic_a = 2.69 Å and c=3.69Å𝑐3.69Åc=3.69\thinspace\textrm{\AA}italic_c = 3.69 Å as the lattice constants. Perdew-Burke-Ernzerhof (PBE) type of generalized gradient approximation (GGA) was used as the exchange-correlation functional[43]. The spectral function A(ϵ,𝐤)𝐴italic-ϵ𝐤A(\epsilon,\mathbf{k})italic_A ( italic_ϵ , bold_k ) along the high-symmetry route is illustrated in Fig. 1(a). This electron spectrum changes significantly when the high-entropy random mixing was slightly turned on for the Fe sites, that is, reducing the probability of the Fe occupation to 0.80.80.80.8 while simultaneously turning on the probability of (CoNiMnCu)0.2subscriptCoNiMnCu0.2(\textrm{CoNiMnCu})_{0.2}( CoNiMnCu ) start_POSTSUBSCRIPT 0.2 end_POSTSUBSCRIPT accordingly.

To fully understand the band broadening, we further performed a similar calculation using projector augmented wave pseudopotential method[44, 45] implemented by Vienna Ab initio Simulation Package (VASP)[46, 47]. The eigenvalues of the Kohn-Sham Hamiltonian are illustrated along the same high symmetry route, as shown in Fig. 1(c). Although there are discrepancies in some details, the Questaal spectrum agrees with the VASP bands generally. Due to the random scattering, although all bands are smeared due to the finite life time, different eigenstates have different levels of broadening. To show this, four representative spectra are highlighted in Figs. 1(a-c). Regions 1 and 2 (red solid boxes) contain Fe-heavy bands, as denoted by the light color in Fig. 1(c). As expected, these bands are broadened significantly after turning on the random mixing on the Fe-occupied sites. Unlike Regions 1 and 2, Regions 3 and 4 contain several bands that are heavily mixed with the Pt orbitals. Specifically, in Region 3, all bands are almost purely given by Pt orbitals, whereas in Region 4, the mixing from Pt orbitals becomes heavy when crossing the A point. In these two regions (red dashed ovals), the band smearing mainly occurs for the Fe-dominated bands, whereas the Pt bands remain sharp. This is consistent with the fact that the body-center sites are uniformly occupied by Pt atoms without any random mixing. After calibrating the band structure, we further calculate the MAE for L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT FePt and compare the results with experiments, as shown in Table 1. We used both PBE-GGA and local density approximation (LDA)[48]. The calculated MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT from the Questaal matches closely with the experimental and the VASP results[49]. In the case of MAE, it appears that DFT overestimates the MAE by roughly an order of magnitude, which is captured by both VASP and Questaal.

Table 1: Comparison of MAE and MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT for pure L10-FePt𝐿subscript10-FePtL1_{0}\textrm{-FePt}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT -FePt.
Method MAE (ergs/cm3superscriptergs/cm3\textrm{ergs/cm}^{3}ergs/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT) MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT (emu/cm3superscriptemu/cm3\textrm{emu/cm}^{3}emu/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT)
Experiment[50, 51] (6.610)×107similar-to6.610superscript107(6.6\sim 10)\times 10^{7}( 6.6 ∼ 10 ) × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT 1100±100plus-or-minus11001001100\pm 1001100 ± 100
VASP (LDA)[49] 1.78×1081.78superscript1081.78\times 10^{8}1.78 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT 1012101210121012
Questaal (LDA) 1.65×1081.65superscript1081.65\times 10^{8}1.65 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT 1074107410741074
VASP (GGA)[49] 1.57×1081.57superscript1081.57\times 10^{8}1.57 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT 1066106610661066
Questaal (GGA) 1.56×1081.56superscript1081.56\times 10^{8}1.56 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT 1097109710971097

Following this verification study, we explore the magnetic properties of high-entropy alloys by fully turning on the random mixing. We first examine the equiatomic FeCoNiMnCu in the FCC phase, and the results are summarized in Fig. 2. The lattice constant of a conventional cell is set to the experimental value a=3.57Å𝑎3.57Åa=3.57\thinspace\textrm{\AA}italic_a = 3.57 Å, which was previously captured by X-ray diffraction[34]. Taking different types of spins for each high-entropy element as separate CPA species, we first examine the ground-state spin configuration of the equiatomic case without considering spin-orbit coupling (SOC). After the self-consistent convergence, positive and negative spins of Mn atoms coexist in the ground state, whereas all other elements are ferromagnetically aligned. This is consistent with existing theoretical results[52]. For the equiatomic composition, LDA gives MS=663emu/cm3subscript𝑀𝑆663superscriptemu/cm3M_{S}=663\thinspace\textrm{emu/cm}^{3}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = 663 emu/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, whereas the GGA value is 705emu/cm3705superscriptemu/cm3705\thinspace\textrm{emu/cm}^{3}705 emu/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. After considering the mass density using the lattice constant, these two values correspond to 81.9emu/g81.9emu/g81.9\thinspace\textrm{emu/g}81.9 emu/g and 87.1emu/g87.1emu/g87.1\thinspace\textrm{emu/g}87.1 emu/g, respectively. These results are close to the DFT (GGA) result (80.8emu/g80.8emu/g80.8\thinspace\textrm{emu/g}80.8 emu/g) using a similar approach[52].

To capture the landscape of MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT near the equiatomic case, the compositional space was explored by varying the concentration of one high-entropy element, whereas simultaneously changing the concentration of others accordingly, kee** the total occupation probability normalized. The LDA results are shown in Fig. 2(b). Here, MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT increases with the Fe and Co concentration, whereas extra mixing of Mn and Cu decreases the MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT significantly. This is expected since Mn is found to have a mixture of both positive and negative spins, whereas Cu atoms are non-magnetic.

Refer to caption
Fig 2: Magnetic properties of the high-entropy crystal FeCoNiMnCu in the FCC phase. (a) The conventional FCC unit cell with all atomic sites randomly occupied by the five different elements. (b) The saturation magnetization when modulating the concentration of one high-entropy element. All other high-entropy elements are adjusted proportionately to maintain a total concentration of 100%percent100100\%100 %. The central composition of 0.20.20.20.2 corresponds to the equiatomic case. (c) The modulation of Curie temperature corresponding to changes in the concentration of one element, according to the pattern described in (b).
Refer to caption
Fig 3: The saturation magnetization and the Curie temperature of high-entropy magnet (FeCoNiMnCu)Pt in the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase. (a) The primitive unit cell of the tetragonal lattice. (b) The saturation magnetization when modulating the concentration of one high-entropy element on the corner sites. All other corner sites are adjusted proportionately to maintain a total concentration of 100%percent100100\%100 %. The body-center Pt site is assumed to be uniformly occupied. The 20%percent2020\%20 % concentration corresponds to the equiatomic case for all the corner sites. (c) The modulation of Curie temperature corresponding to changes in the concentration of one element, following the pattern described in (b).

After exploring the magnetic ordering of equiatomic recipe, we estimate the ordering temperature (TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT) through mean-field approximation.[53, 54, 55]:

TC=2(ϵPMϵGS)3(1c)kB,subscript𝑇𝐶2subscriptitalic-ϵPMsubscriptitalic-ϵGS31𝑐subscript𝑘BT_{C}=\frac{2(\epsilon_{\textrm{PM}}-\epsilon_{\textrm{GS}})}{3(1-c)k_{\textrm% {B}}},italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = divide start_ARG 2 ( italic_ϵ start_POSTSUBSCRIPT PM end_POSTSUBSCRIPT - italic_ϵ start_POSTSUBSCRIPT GS end_POSTSUBSCRIPT ) end_ARG start_ARG 3 ( 1 - italic_c ) italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT end_ARG , (1)

where ϵPMsubscriptitalic-ϵPM\epsilon_{\textrm{PM}}italic_ϵ start_POSTSUBSCRIPT PM end_POSTSUBSCRIPT and ϵGSsubscriptitalic-ϵGS\epsilon_{\textrm{GS}}italic_ϵ start_POSTSUBSCRIPT GS end_POSTSUBSCRIPT represent converged energies of paramagnetic and ground-state spin configurations, respectively. Here, kBsubscript𝑘Bk_{\textrm{B}}italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT is the Boltzmann constant and c𝑐citalic_c denotes the net concentration of the non-magnetic atoms in the HEA. Note that Eq. 1 is a crude approximation for estimating the upper limit[53]. The estimated TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT for the equiatomic case using LDA and GGA are found to be 261K261K261\thinspace\textrm{K}261 K and 359K359K359\thinspace\textrm{K}359 K, respectively, which are close to existing DFT results[52]. We then explore the landscape of TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT in the compositional space near the equiatomic recipe by varying the concentrations similar to that used for MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT. The LDA results are summarized in Fig. 2(c), where TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT decreases significantly with the increase in the Mn concentration. This is consistent with the fact that the ground-state Mn atoms have a mixing of both positive and negative spins, suggesting that their exchange couplings with other high-entropy species should be antiferromagnetic.

To understand the MAE of FeCoNiMnCu in its FCC solid solution phase, we compare the ground-state energy when the spin quantization axis is along different directions. Care must be taken since the MAE is only a small fraction of the total energy. To evaluate the limit of our numerical resolution, we first examine the energy difference between two equivalent directions within the 100delimited-⟨⟩100\langle 100\rangle⟨ 100 ⟩ family. With a convergence criterion of 106Rydsuperscript106Ryd10^{-6}\thinspace\textrm{Ryd}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT Ryd, the energy difference between [001]delimited-[]001[001][ 001 ] and [100]delimited-[]100[100][ 100 ] corresponds to an MAE of 1.82×106ergs/cm31.82superscript106superscriptergs/cm31.82\times 10^{6}\thinspace\textrm{ergs/cm}^{3}1.82 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ergs/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. Although this energy scale is sizable and can indeed be detected experimentally[34], it is difficult for DFT calculations to resolve energies that are smaller or near this range. We further examine the MAE value when taking [111]delimited-[]111[111][ 111 ] and [110]delimited-[]110[110][ 110 ] as the easy and hard axes, respectively, which resulted in an MAE of 2.01×106ergs/cm32.01superscript106superscriptergs/cm32.01\times 10^{6}\thinspace\textrm{ergs/cm}^{3}2.01 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ergs/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. This is close to the resolution of our calculation, and we therefore conclude that the MAE between [111]delimited-[]111[111][ 111 ] and [110]delimited-[]110[110][ 110 ] is beyond the numerical accuracy. Such a small MAE is a consequence of the highly symmetric cubical lattice, as well as the lack of heavy elements providing a sizable SOC. To increase the MAE, we further mix Pt into the FCC solid solution by replacing Mn[56], which is known to couple to other high-entropy species antiferromagnetically. Kee** the lattice constant the same to the case of FeCoNiCuMn, the MAE with the Pt mixing remains close to our numerical resolution (106ergs/cm3similar-toabsentsuperscript106superscriptergs/cm3\sim 10^{6}\textrm{ergs/cm}^{3}∼ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ergs/cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT). This suggests that further breaking the symmetries of the cubical lattice may be necessary to increase the MAE.

One way to break the symmetry of the FCC lattice is to insert a layer of Pt atoms between the high-entropy layers stacking along the c𝑐citalic_c axis. The change in the c𝑐citalic_c axis brings the FCC lattice to a tetragonal one, as shown by Fig. 3(a). Pure crystals of this phase is well known to host large MAE in many different cases such as FePt, CoPt and MnAl[57, 58, 59]. Recently, the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase of (FeCoNiMnCu)Pt has been experimentally realized through a co-sputtering followed by a rapid thermal annealing, where both good crystallization and large MAE have been confirmed[32, 60, 34]. According to the experimental characterization, we assume that there is no random mixing for the Pt atoms at the body-center sites, whereas all corner sites are occupied by the high-entropy mixing of the transition metal elements. This assures that the strong SOC given by the Pt atoms is homogeneously experienced by all the transition metal elements. Using the experimental lattice constants listed in Fig. 3(a), we explore the magnetic properties of (FeCoNiMnCu)Pt in the compositional space, and the results are shown in Figs. 3(b-c). Compared to the FCC case, the range of MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT becomes smaller near the equiatomic composition due to the insertion of the non-magnetic Pt. Unlike the equiatomic FCC case, all compositions we explored for the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase have a ferromagnetic ground state, resulting in a significant increase of 200Ksimilar-toabsent200K\sim 200\thinspace\textrm{K}∼ 200 K in TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. Increasing the concentrations of Fe and Co increases both MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT and TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, whereas the mixing of Mn and Cu has an opposite trend. This suggests a potential engineering strategy to reduce the writing power for HAMR devices.

We use machine learning to explore the MAE of the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT phase, resulting in a smooth profile in the 5superscript5\mathbb{R}^{5}blackboard_R start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT compositional space that is consistent with the experimental trend. Similar to the previous cases, we first modulate the compositions near the equiatomic case for the transition metals, while kee** Pt fixed at 50%percent5050\%50 %. For all explored cases, DFT suggests that the [001]delimited-[]001[001][ 001 ] direction is an easier axis compared to [100]delimited-[]100[100][ 100 ], and the MAE values are shown in Fig. 4(a). Although this result suggests significant correlation between the Fe concentration and the MAE, it is clear that all transition-metal elements have strong influences. To fully capture and visualize the MAE profile, we use machine learning and interpolation in the compositional space. This is implemented using multidimensional scaling (MDS), an unsupervised learning method typically used to cluster high-dimensional points [61]. MDS searches for an optimal curved 2-dimensional surface in a high (5 in our case) dimensional parameter space, such that the points can uniformly scale their Euclidean distance to their neighbors after projected onto the surface. This surface is numerically described by a discrete manifold, which can be illustrated on a flat 2D plane after a straightforward transformation.

Refer to caption
Fig 4: The magnetic anisotropy energy density (MAE) of the L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT high-entropy magnet (FeCoNiMnCu)Pt. (a) The variation of MAE for different compositions when each high entropy element is modulated following the pattern described in Figs. 2 and 3. (b) The compositions in (a) projected from 5superscript5\mathbb{R}^{5}blackboard_R start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT to 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT using Multidimensional Scaling (MDS), which preserves the Euclidean distances between points. Color saturation denotes the MAE values scaled between 00 and 1111. The red and blue points denote the DFT and the experimental results, respectively. (c-d) Direct comparison between the theoretical and experimental results for the 5555 samples highlighted in (b). Here the MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT comparison is shown in (c), whereas the MAE comparison is shown in (d). Note that MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT is compared with the same scale in (c), whereas the MAE comparison is illustrated on different scales in (d).

The flattened MDS map of the data points in Fig.4(a) is illustrated in Fig.4(b), with the projected locations denoted by the crosses. The MAE values obtained by DFT are illustrated with the white-salmon color scale, where the empty space between points are filled using Hermite-Zero-Derivative Start (HZDS) interpolation. The DFT results are compared with experiments using five different L10𝐿subscript10L1_{0}italic_L 1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT thin films near the equiatomic composition. These samples were fabricated and characterized using the same approach described in our previous work[34], where the compositions are estimated by Energy-Dispersive X-Ray Spectroscopy (EDS) and the MAE values estimated by magnetometry measurements. The specific compositions of these five samples are listed in Table 2.

Table 2: Elemental compositions of the samples.
Samples Compositions
1 Fe0.07Co0.12Ni0.11Mn0.11Cu0.12Pt0.47subscriptFe0.07subscriptCo0.12subscriptNi0.11subscriptMn0.11subscriptCu0.12subscriptPt0.47\textrm{Fe}_{0.07}\textrm{Co}_{0.12}\textrm{Ni}_{0.11}\textrm{Mn}_{0.11}% \textrm{Cu}_{0.12}\textrm{Pt}_{0.47}Fe start_POSTSUBSCRIPT 0.07 end_POSTSUBSCRIPT Co start_POSTSUBSCRIPT 0.12 end_POSTSUBSCRIPT Ni start_POSTSUBSCRIPT 0.11 end_POSTSUBSCRIPT Mn start_POSTSUBSCRIPT 0.11 end_POSTSUBSCRIPT Cu start_POSTSUBSCRIPT 0.12 end_POSTSUBSCRIPT Pt start_POSTSUBSCRIPT 0.47 end_POSTSUBSCRIPT
2 Fe0.10Co0.08Ni0.20Mn0.07Cu0.09Pt0.46subscriptFe0.10subscriptCo0.08subscriptNi0.20subscriptMn0.07subscriptCu0.09subscriptPt0.46\textrm{Fe}_{0.10}\textrm{Co}_{0.08}\textrm{Ni}_{0.20}\textrm{Mn}_{0.07}% \textrm{Cu}_{0.09}\textrm{Pt}_{0.46}Fe start_POSTSUBSCRIPT 0.10 end_POSTSUBSCRIPT Co start_POSTSUBSCRIPT 0.08 end_POSTSUBSCRIPT Ni start_POSTSUBSCRIPT 0.20 end_POSTSUBSCRIPT Mn start_POSTSUBSCRIPT 0.07 end_POSTSUBSCRIPT Cu start_POSTSUBSCRIPT 0.09 end_POSTSUBSCRIPT Pt start_POSTSUBSCRIPT 0.46 end_POSTSUBSCRIPT
3 Fe0.11Co0.09Ni0.10Mn0.09Cu0.11Pt0.50subscriptFe0.11subscriptCo0.09subscriptNi0.10subscriptMn0.09subscriptCu0.11subscriptPt0.50\textrm{Fe}_{0.11}\textrm{Co}_{0.09}\textrm{Ni}_{0.10}\textrm{Mn}_{0.09}% \textrm{Cu}_{0.11}\textrm{Pt}_{0.50}Fe start_POSTSUBSCRIPT 0.11 end_POSTSUBSCRIPT Co start_POSTSUBSCRIPT 0.09 end_POSTSUBSCRIPT Ni start_POSTSUBSCRIPT 0.10 end_POSTSUBSCRIPT Mn start_POSTSUBSCRIPT 0.09 end_POSTSUBSCRIPT Cu start_POSTSUBSCRIPT 0.11 end_POSTSUBSCRIPT Pt start_POSTSUBSCRIPT 0.50 end_POSTSUBSCRIPT
4 Fe0.13Co0.11Ni0.07Mn0.09Cu0.11Pt0.49subscriptFe0.13subscriptCo0.11subscriptNi0.07subscriptMn0.09subscriptCu0.11subscriptPt0.49\textrm{Fe}_{0.13}\textrm{Co}_{0.11}\textrm{Ni}_{0.07}\textrm{Mn}_{0.09}% \textrm{Cu}_{0.11}\textrm{Pt}_{0.49}Fe start_POSTSUBSCRIPT 0.13 end_POSTSUBSCRIPT Co start_POSTSUBSCRIPT 0.11 end_POSTSUBSCRIPT Ni start_POSTSUBSCRIPT 0.07 end_POSTSUBSCRIPT Mn start_POSTSUBSCRIPT 0.09 end_POSTSUBSCRIPT Cu start_POSTSUBSCRIPT 0.11 end_POSTSUBSCRIPT Pt start_POSTSUBSCRIPT 0.49 end_POSTSUBSCRIPT
5 Fe0.19Co0.10Ni0.10Mn0.07Cu0.08Pt0.46subscriptFe0.19subscriptCo0.10subscriptNi0.10subscriptMn0.07subscriptCu0.08subscriptPt0.46\textrm{Fe}_{0.19}\textrm{Co}_{0.10}\textrm{Ni}_{0.10}\textrm{Mn}_{0.07}% \textrm{Cu}_{0.08}\textrm{Pt}_{0.46}Fe start_POSTSUBSCRIPT 0.19 end_POSTSUBSCRIPT Co start_POSTSUBSCRIPT 0.10 end_POSTSUBSCRIPT Ni start_POSTSUBSCRIPT 0.10 end_POSTSUBSCRIPT Mn start_POSTSUBSCRIPT 0.07 end_POSTSUBSCRIPT Cu start_POSTSUBSCRIPT 0.08 end_POSTSUBSCRIPT Pt start_POSTSUBSCRIPT 0.46 end_POSTSUBSCRIPT

These compositions are projected to the same MDS manifold, which are illustrated by the solid circles in Fig. 4(b) where the white-blue scale represents the normalized experimental values. It can be seen that the general trend of the experimental MAE is highly correlated to the interpolated MDS profile. We also performed DFT calculations using the experimental compositions of the five samples, and the results of MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT and MAE are illustrated in Figs. 4(c-d). In both cases, DFT overestimates the values by at least several factors, which is consistent with the trend seen in the case of pure FePt (Table 1). This may be induced by the assumption of perfect crystallization, whereas the real samples have complicated grain landscapes and lattice distortions that are difficult to capture by the DFT setup. The overestimation is also likely induced by the lost information of the mean-field approach in CPA, where consecutive scattering is ignored. Although quantitative agreement between theory and experiment is difficult, the trends of MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT and MAE are both consistent with the experimental ones. Note that the experimental compositions are not used in the interpolation shown in Fig. 4 (b), suggesting that the smooth profile of the MAE can be captured by the sparse DFT points sampled in the compositional space.

The results shown above suggest that sparse sampling in the compositional space of HEAs using primitive-cell DFT calculations can capture the trend of magnetic properties including MSsubscript𝑀𝑆M_{S}italic_M start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT, MAE and TCsubscript𝑇𝐶T_{C}italic_T start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, even after losing the information of atomic occupation details within CPA. Since CPA assumes translational symmetry, the Green’s functions are naturally diagonal in 𝐤𝐤\mathbf{k}bold_k, allowing for the fast evaluation of the electron density by working only within the first Brillouin zone. Although the experiments do not quantitatively agree with DFT, the profile in the high-dimensional compositional space is smooth, which can be captured by a reasonable number of discrete points sparsely sampled in the compositional space. This invites further research to establish large-scale databases within the mean-field approximation of CPA to establish predictive machine-learning models. This also introduces an intriguing opportunity to achieve the fast design of future high-entropy magnets for many scenarios of application.

Acknowledgments: This paper is based upon work supported by the National Science Foundation (US) under Grant No. ECCS-2151809. This work used Bridges-2 at Pittsburgh Supercomputing Center through allocation PHY230018 from the Advanced Cyberinfrastructure Coordination Ecosystem: Services & Support (ACCESS) program, which is supported by National Science Foundation (US) grants #2138259, #2138286, #2138307, #2137603, and #2138296. The acquisition of the MPMS3 system used in this study was supported by the NSF-MRI program (DMR-1828420). TS and SNK gratefully acknowledge funding from US Department of Energy (DOE), under the award DE-SC0006420. JJ acknowledges support under the CCP9 project "Computational Electronic Structure of Condensed Matter"(part of Computational Science Centre of Research Communities(CoSeC)).

References

  • Zhang et al. [2014] Y. Zhang, T. T. Zuo, Z. Tang, M. C. Gao, K. A. Dahmen, P. K. Liaw, and Z. P. Lu, Microstructures and properties of high-entropy alloys, Prog. Mater. Sci. 61, 1 (2014).
  • Yeh et al. [2004a] J.-W. Yeh, S.-J. Lin, T.-S. Chin, J.-Y. Gan, S.-K. Chen, T.-T. Shun, C.-H. Tsau, and S.-Y. Chou, Formation of simple crystal structures in Cu-Co-Ni-Cr-Al-Fe-Ti-V alloys with multiprincipal metallic elements, Met. Mater Trans A 35, 2533 (2004a).
  • Chen et al. [2004] T. K. Chen, T. T. Shun, J. W. Yeh, and M. S. Wong, Nanostructured nitride films of multi-element high-entropy alloys by reactive DC sputtering, Surf. Coat. Tech. 188–189, 193 (2004).
  • Cantor et al. [2004] B. Cantor, I. T. H. Chang, P. Knight, and A. J. B. Vincent, Microstructural development in equiatomic multicomponent alloys, Mater. Sci. Eng. A 375–377, 213 (2004).
  • Yeh et al. [2004b] J.-W. Yeh, S.-K. Chen, S.-J. Lin, J.-Y. Gan, T.-S. Chin, T.-T. Shun, C.-H. Tsau, and S.-Y. Chang, Nanostructured High-Entropy Alloys with Multiple Principal Elements: Novel Alloy Design Concepts and Outcomes, Adv. Eng. Mater. 6, 299 (2004b).
  • Miracle and Senkov [2017] D. B. Miracle and O. N. Senkov, A critical review of high entropy alloys and related concepts, Acta Mater. 122, 448 (2017).
  • George et al. [2020] E. P. George, W. A. Curtin, and C. C. Tasan, High entropy alloys: A focused review of mechanical properties and deformation mechanisms, Acta Mater. 188, 435 (2020).
  • George et al. [2019] E. P. George, D. Raabe, and R. O. Ritchie, High-entropy alloys, Nat. Rev. Mater. 4, 515 (2019).
  • Kozak et al. [2015] R. Kozak, A. Sologubenko, and W. Steurer, Single-phase high-entropy alloys – an overview, Z Krist. Cryst Mater 230, 55 (2015).
  • Praveen and Kim [2018] S. Praveen and H. S. Kim, High-Entropy Alloys: Potential Candidates for High-Temperature Applications - An Overview, Adv. Eng. Mater. 20, 1700645 (2018).
  • Lei et al. [2018] Z. Lei, X. Liu, Y. Wu, H. Wang, S. Jiang, S. Wang, X. Hui, Y. Wu, B. Gault, P. Kontis, D. Raabe, L. Gu, Q. Zhang, H. Chen, H. Wang, J. Liu, K. An, Q. Zeng, T.-G. Nieh, and Z. Lu, Enhanced strength and ductility in a high-entropy alloy via ordered oxygen complexes, Nature 563, 546 (2018).
  • Li et al. [2021] W. Li, D. Xie, D. Li, Y. Zhang, Y. Gao, and P. K. Liaw, Mechanical behavior of high-entropy alloys, Prog. Mater. Sci. 118, 100777 (2021).
  • Chen et al. [2005] Y. Y. Chen, T. Duval, U. D. Hung, J. W. Yeh, and H. C. Shih, Microstructure and electrochemical properties of high entropy alloys—a comparison with type-304 stainless steel, Corros. Sci. 47, 2257 (2005).
  • Lee et al. [2007] C. P. Lee, Y. Y. Chen, C. Y. Hsu, J. W. Yeh, and H. C. Shih, The Effect of Boron on the Corrosion Resistance of the High Entropy Alloys Al0.5CoCrCuFeNiBx, J. Electrochem. Soc. 154, C424 (2007).
  • Law and Franco [2023] J. Y. Law and V. Franco, Review on magnetocaloric high-entropy alloys: Design and analysis methods, J. Mater. Res. 38, 37 (2023).
  • Vaidya et al. [2019] M. Vaidya, G. M. Muralikrishna, and B. S. Murty, High-entropy alloys by mechanical alloying: A review, J. Mater. Res. 34, 664 (2019).
  • Ye et al. [2016] Y. F. Ye, Q. Wang, J. Lu, C. T. Liu, and Y. Yang, High-entropy alloy: Challenges and prospects, Mater. Today 19, 349 (2016).
  • Rao et al. [2022] Z. Rao, P.-Y. Tung, R. Xie, Y. Wei, H. Zhang, A. Ferrari, T. Klaver, F. Körmann, P. T. Sukumar, A. Kwiatkowski da Silva, Y. Chen, Z. Li, D. Ponge, J. Neugebauer, O. Gutfleisch, S. Bauer, and D. Raabe, Machine learning–enabled high-entropy alloy discovery, Science 378, 78 (2022).
  • Divilov et al. [2024] S. Divilov, H. Eckert, D. Hicks, C. Oses, C. Toher, R. Friedrich, M. Esters, M. J. Mehl, A. C. Zettel, Y. Lederer, E. Zurek, J.-P. Maria, D. W. Brenner, X. Campilongo, S. Filipović, W. G. Fahrenholtz, C. J. Ryan, C. M. DeSalle, R. J. Crealese, D. E. Wolfe, A. Calzolari, and S. Curtarolo, Disordered enthalpy–entropy descriptor for high-entropy ceramics discovery, Nature 625, 66 (2024).
  • Huang et al. [2019] W. Huang, P. Martin, and H. L. Zhuang, Machine-learning phase prediction of high-entropy alloys, Acta Mater. 169, 225 (2019).
  • Oses et al. [2018] C. Oses, E. Gossett, D. Hicks, F. Rose, M. J. Mehl, E. Perim, I. Takeuchi, S. Sanvito, M. Scheffler, Y. Lederer, O. Levy, C. Toher, and S. Curtarolo, AFLOW-CHULL: Cloud-Oriented Platform for Autonomous Phase Stability Analysis, J. Chem. Inf. Model. 58, 2477 (2018).
  • Wang et al. [2014] J. Wang, Z. Zheng, J. Xu, and Y. Wang, Microstructure and magnetic properties of mechanically alloyed FeSiBAlNi (Nb) high entropy alloys, J. Magn. Magn. Mater. 355, 58 (2014).
  • Liu et al. [2012] L. Liu, J. B. Zhu, J. C. Li, and Q. Jiang, Microstructure and Magnetic Properties of FeNiCuMnTiSnx High Entropy Alloys, Adv. Eng. Mater. 14, 919 (2012).
  • Tang et al. [2021] N. Tang, L. Quigley, W. L. Boldman, C. S. Jorgensen, R. Koch, D. O’Leary, H. R. Medal, P. D. Rack, and D. A. Gilbert, Magnetism in metastable and annealed compositionally complex alloys, Phys. Rev. Mater. 5, 114405 (2021).
  • Na et al. [2021] S.-M. Na, P. K. Lambert, and N. J. Jones, Hard magnetic properties of FeCoNiAlCuXTiX based high entropy alloys, AIP Adv. 11, 015210 (2021).
  • Kumari et al. [2022] P. Kumari, A. K. Gupta, R. K. Mishra, M. Ahmad, and R. R. Shahi, A Comprehensive Review: Recent Progress on Magnetic High Entropy Alloys and Oxides, J. Magn. Magn. Mater. 554, 169142 (2022).
  • Weller and Moser [1999] D. Weller and A. Moser, Thermal effect limits in ultrahigh-density magnetic recording, IEEE Trans. Magn. 35, 4423 (1999).
  • Rottmayer et al. [2006] R. Rottmayer, S. Batra, D. Buechel, W. Challener, J. Hohlfeld, Y. Kubota, L. Li, B. Lu, C. Mihalcea, K. Mountfield, K. Pelhos, C. Peng, T. Rausch, M. Seigler, D. Weller, and X.-M. Yang, Heat-Assisted Magnetic Recording, IEEE Trans. Magn. 42, 2417 (2006).
  • Weller et al. [2016] D. Weller, G. Parker, O. Mosendz, A. Lyberatos, D. Mitin, N. Y. Safonova, and M. Albrecht, Review Article: FePt heat assisted magnetic recording media, J. Vac. Sci. Technol. B 34, 060801 (2016).
  • Davies et al. [2004] J. E. Davies, O. Hellwig, E. E. Fullerton, G. Denbeaux, J. B. Kortright, and K. Liu, Magnetization reversal of Co/Pt multilayers: Microscopic origin of high-field magnetic irreversibility, Phys. Rev. B 70, 224434 (2004).
  • Rahman et al. [2009] M. T. Rahman, R. K. Dumas, N. Eibagi, N. N. Shams, Y.-C. Wu, K. Liu, and C.-H. Lai, Controlling magnetization reversal in Co/Pt nanostructures with perpendicular anisotropy, Appl. Phys. Lett. 94, 042507 (2009).
  • Gilbert et al. [2013] D. A. Gilbert, L.-W. Wang, T. J. Klemmer, J.-U. Thiele, C.-H. Lai, and K. Liu, Tuning magnetic anisotropy in (001) oriented L10 (Fe1-xCux)55Pt45 films, Appl. Phys. Lett. 102, 132406 (2013).
  • Cuadrado et al. [2016] R. Cuadrado, K. Liu, T. J. Klemmer, and R. W. Chantrell, In-plane/out-of-plane disorder influence on the magnetic anisotropy of Fe1-yMnyPt-L10 bulk alloy, Appl. Phys. Lett. 108, 123102 (2016).
  • Beeson et al. [2023] W. B. Beeson, D. Bista, H. Zhang, S. Krylyuk, A. Davydov, G. Yin, and K. Liu, Single-phase L10subscriptL10\textrm{L1}_{0}L1 start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT-ordered high entropy thin films with high magnetic anisotropy, arXiv:2311.06618 (2023).
  • Soven [1967] P. Soven, Coherent-Potential Model of Substitutional Disordered Alloys, Phys. Rev. 156, 809 (1967).
  • Soven [1969] P. Soven, Contribution to the Theory of Disordered Alloys, Phys. Rev. 178, 1136 (1969).
  • Woodgate and Staunton [2022] C. D. Woodgate and J. B. Staunton, Compositional phase stability in medium-entropy and high-entropy Cantor-Wu alloys from an ab initio all-electron Landau-type theory and atomistic modeling, Phys. Rev. B 105, 115124 (2022).
  • Grimme et al. [2010] S. Grimme, J. Antony, S. Ehrlich, and H. Krieg, A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys. 132, 154104 (2010).
  • Becke and Johnson [2005] A. D. Becke and E. R. Johnson, A density-functional model of the dispersion interaction, J. Chem. Phys. 123, 154101 (2005).
  • Grimme et al. [2011] S. Grimme, S. Ehrlich, and L. Goerigk, Effect of the dam** function in dispersion corrected density functional theory, J. Comput. Chem. 32, 1456 (2011).
  • Pashov et al. [2020] D. Pashov, S. Acharya, W. R. Lambrecht, J. Jackson, K. D. Belashchenko, A. Chantis, F. Jamet, and M. van Schilfgaarde, Questaal: A package of electronic structure methods based on the linear muffin-tin orbital technique, Comput. Phys. Commun. 249, 107065 (2020).
  • Monkhorst and Pack [1976] H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B 13, 5188 (1976).
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett. 77, 3865 (1996).
  • Blöchl [1994] P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50, 17953 (1994).
  • Kresse and Joubert [1999] G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B 59, 1758 (1999).
  • Kresse and Furthmüller [1996] G. Kresse and J. Furthmüller, Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci. 6, 15 (1996).
  • Kresse and Furthmüller [1996] G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54, 11169 (1996).
  • von Barth and Hedin [1972] U. von Barth and L. Hedin, A local exchange-correlation potential for the spin polarized case. I, J. Phys. C: Solid State Phys. 5, 1629 (1972).
  • Wolloch et al. [2017] M. Wolloch, D. Suess, and P. Mohn, Influence of antisite defects and stacking faults on the magnetocrystalline anisotropy of FePt, Phys. Rev. B 96, 104408 (2017).
  • Klemmer et al. [1995] T. Klemmer, D. Hoydick, H. Okumura, B. Zhang, and W. A. Soffa, Magnetic hardening and coercivity mechanisms in L10 ordered FePd ferromagnets, Scr. Mater. 33, 1793 (1995).
  • Weller et al. [2000] D. Weller, A. Moser, L. Folks, M. Best, W. Lee, M. Toney, M. Schwickert, J.-U. Thiele, and M. Doerner, High Ku materials approach to 100 Gbits/in2, IEEE Trans. Magn. 36, 10 (2000).
  • Rao et al. [2020] Z. Rao, B. Dutta, F. Körmann, D. Ponge, L. Li, J. He, L. Stephenson, L. Schäfer, K. Skokov, O. Gutfleisch, D. Raabe, and Z. Li, Unveiling the mechanism of abnormal magnetic behavior of FeNiCoMnCu high-entropy alloys through a joint experimental-theoretical study, Phys. Rev. Mater. 4, 014402 (2020).
  • Sato et al. [2010] K. Sato, L. Bergqvist, J. Kudrnovský, P. H. Dederichs, O. Eriksson, I. Turek, B. Sanyal, G. Bouzerar, H. Katayama-Yoshida, V. A. Dinh, T. Fukushima, H. Kizaki, and R. Zeller, First-principles theory of dilute magnetic semiconductors, Rev. Mod. Phys. 82, 1633 (2010).
  • Körmann et al. [2015] F. Körmann, D. Ma, D. D. Belyea, M. S. Lucas, C. W. Miller, B. Grabowski, and M. H. F. Sluiter, “Treasure maps” for magnetic high-entropy-alloys from theory and experiment, Appl. Phys. Lett. 107, 142404 (2015).
  • Sato et al. [2003] K. Sato, P. H. Dederics, and H. Katayama-Yoshida, Curie temperatures of III–V diluted magnetic semiconductors calculated from first principles, Europhys. Lett. 61, 403 (2003).
  • Kurniawan et al. [2016] M. Kurniawan, A. Perrin, P. Xu, V. Keylin, and M. McHenry, Curie Temperature Engineering in High Entropy Alloys for Magnetocaloric Applications, IEEE Magn. Lett. 7, 1 (2016).
  • Weller et al. [1992] D. Weller, H. Brändle, G. Gorman, C.-J. Lin, and H. Notarys, Magnetic and magneto-optical properties of cobalt-platinum alloys with perpendicular magnetic anisotropy, Appl. Phys. Lett. 61, 2726 (1992).
  • Sakuma [1994] A. Sakuma, Electronic Structure and Magnetocrystalline Anisotropy Energy of MnAl, J. Phys. Soc. Jpn. 63, 1422 (1994).
  • Lyubina et al. [2005] J. Lyubina, I. Opahle, K.-H. Müller, O. Gutfleisch, M. Richter, M. Wolf, and L. Schultz, Magnetocrystalline anisotropy in L10 FePt and exchange coupling in FePt/Fe3Pt nanocomposites, J. Phys. Condens. Matter 17, 4157 (2005).
  • Gilbert et al. [2014] D. A. Gilbert, J.-W. Liao, L.-W. Wang, J. W. Lau, T. J. Klemmer, J.-U. Thiele, C.-H. Lai, and K. Liu, Probing the A1 to L10 transformation in FeCuPt using the first order reversal curve method, APL Mater. 2, 086106 (2014).
  • Kruskal [1964] J. B. Kruskal, Nonmetric multidimensional scaling: A numerical method, Psychometrika 29, 115 (1964).