We investigate the correlation between geometrical information, stability, and magnetization of SmFe12-based structures using machine learning-aided genetic algorithm structure generation and first-principle calculation. In parallel with structure generation inherited using the USPEX program, a pool of structures is created for every population using the sub-symmetry perturbation method. A framework using embedded orbital field matrix representation as structure fingerprint and Gaussian process as a predictor has been applied to ranking the most potential stability structures. As a result, the original structure SmFe12 with the well-known tetragonal I4/mmm symmetry is investigated with a parabolic dependence between formation energy and its magnetization by continuous distortions of the unit-cell lattice parameter and individual sites. Notably, a SmFe12 structure with I4/mmm symmetry is found with 7.5% increasing magnetization while keeping the similar formation energy with the most stable structures in this family. With SmFe11CoN family, structures with N interstitial position in the center of Sm and Fe octahedron show outperform all other structures in both ability of stabilization and remaining high magnetization of the original structure. Finally, further investigation using metric learning embedding space brings valuable insight into the correlation between geometrical arrangement, stability, and magnetization of this structure family.

Despite the lively discussion in the community,1–4 density functional theory is still the most popular approach for calculating the electronic structure of molecules and extended materials.5,6 When the crystal structure is known, many properties can be computed reliably using a first-principle calculation, such as volumes within 1%–3% (LDA, GGA), bandgaps within 30% underestimated (LDA, GGA), and 10% (GW) and somehow unsatisfactory—for van der Waals crystals or systems with localized d- and f-electrons. However, when we do not know detail structure information of the compounds or its thermal decomposition mechanism, finding new stable and outperforming one seemly as infeasible was summarized by the one-word paper by Gavezzotti.7 Therefore, it is necessary to investigate the correlation between its physical properties, e.g., stability, as the function of geometrical information, including element site properties and lattice symmetry configuration. We introduce machine learning (ML) based method approaching to this problem and taking the SmFe12 compound as the object for investigation.

The SmFe12 structure with tetragonal I4/mmm symmetry (space group 139) for the ThMn12 prototype structure is expected as the next generation magnet with outperform magnetic properties as high saturation magnetization, magnetocrystalline anisotropy, and Curie temperature.8 However, SmFe12 bulk crystal have not been synthesized due to the difficulty of stabilizing the material in nature. Modifying element site properties such as substituting a part of the Fe atom with the other atom in the SmFe12 structure may derive a bundle of structures in which the difference relies on the occupied position of the new atom. Many first-principle studies have been conducted by replacing Co, Ti, V, Cr, Mo, W, or Ga to the original Fe site of SmFe12 or replacing Zr atom to rare-earth sites9,10 to obtain a stable ThMn12-type phase.11–17 Unrestricted from ternary compounds, recent researchers have emphasized searching for the most potentially stable SmFe12-based quaternary compounds using the bi-element substitution method.9,10,18–24 In Ref. 18, authors report more than 3000 substituted structures generated by substituting up to three Fe sites of the original structure by other atoms. On the other hand, in the attempt to produce magnets, the difficulty comes from controlling correlations between synthesized parameters, e.g., temperature, pressure, diffusion rate, and statistical information of macro property, e.g., the particle grain size.25–28 The continuous dependence of first-principle calculation properties on the synthesized parameters or the unit-cell geometry could be a promising guideline to accelerate the synthesizing process.

Taking into account perturbation of the lattice parameter as expanding, collapsing the unit cell, and changing its symmetry may increase the number of combinations that may appear in nature. As shown in Ref. 29, the similar ThMn12-type structures, YFe12 with two monoclinic C2/m (space group 12) phases, rather than the original I4/mmm symmetry, were reported in a metastable state by permutating rare-earth site and 8i iron site. These two phases are estimated their total magnetization outperform 4.5% than the original I4/mmm symmetry. There is also an interstitial approach that may also stabilize the desired structure. In Ref. 30, authors reported properties of thin films fabricated SmFe12 and SmFe12Nx with N substituted components. The nitrogenation process was reported to increase the Curie temperature, saturation magnetization, and anisotropy change for all SmFe12 and Sm2Fe17 structures. The number of hypothetical structures will rise significantly to the arbitrary position of the newly added atom.

In this research, we introduce a machine learning-aided genetic algorithm method for efficiently studying SmFe12 compounds aiming to optimize three criteria simultaneously: (1) diversifying the symmetry of screening structure, (2) finding optimal stability, and (3) revealing smoothness correlation between the stability and geometrical information. Descriptions about the machine learning method, genetic algorithm and first-principle calculation are shown in Sec. II. In Sec. III, we present our finding in investigating the structure–property relationship of SmFe12, Co substituted SmFe12, and N interstitial SmFe11Co structures using the proposed method.

We introduce a machine learning-aided genetic algorithm in the structure discovery method consisting of five consecutive stages: (i) structure generator based on symmetry transformation operator, (ii) machine learning recommendation system, (iii) genetic algorithm structure generator, and (iv) first-principle calculation to estimate physical properties. Details illustration of the method is summarized in Fig. 1.

FIG. 1.

(a) Workflow of the machine learning-aided genetic algorithm method in structure discovery. (b) Lattice of maximal subgroups from I4/mmm (139) to Immm (71), created by Bilbao Crystallographic. The maximal subgroups with their corresponding indexes are shown in the brackets. (c) Skeleton of SmFe12 structures with I4/mmm (139), P4/mmm (123), Immm (71), and C2 (5) symmetries. Surfaces connected by Fe atoms are shown to visualize distortions among sub-symmetry transformations better.

FIG. 1.

(a) Workflow of the machine learning-aided genetic algorithm method in structure discovery. (b) Lattice of maximal subgroups from I4/mmm (139) to Immm (71), created by Bilbao Crystallographic. The maximal subgroups with their corresponding indexes are shown in the brackets. (c) Skeleton of SmFe12 structures with I4/mmm (139), P4/mmm (123), Immm (71), and C2 (5) symmetries. Surfaces connected by Fe atoms are shown to visualize distortions among sub-symmetry transformations better.

Close modal

Let us assume, we start with an interested symmetry group G, e.g., I4/mmm (139), in the original structure of SmFe12 with the ThMn12 prototype structure. In this structure, there are four Wyckoff positions with 2a-occupied by Sm atom and 8i, 8j, and 8f occupied by Fe atoms. A new structure with sub-symmetry group H with H<G is created by splitting Wyckoff positions from the original structure with a predefined distortion constant. A simple transformation of the SmFe12 structure with the detailed movement of Fe sites in I4/mmm (139) to Immm (71), and C2/m (12) is shown in Sec. III and distortion of Fe network in Sm local is shown in Fig. 1(c). In this work, we use PyXtal library37 for symmetry transformation operators we applied in the structure generator.

In Fig. 1(b), the graph of possible symmetry transformation beginning with the I4/mmm (139) symmetry is shown, limiting up to the Immm (71) symmetry. Other nodes in the middle correspond to the intermediate subgroups between these symmetries. Two nodes in the graph are connected with an edge if there is a group–subgroup relation between the corresponding subgroups. A node is connected with itself if this symmetry group has an isomorphic subgroup. The graph is created by querying information from Bilbao Crystallographic Server.38–40 More details about sub-symmetry transformation can be found in this Bilbao Crystallographic Server. Figure 1(c) shows example of skeleton of SmFe12 structures with I4/mmm (139), P4/mmm (123), Immm (71), and C2 (5) symmetry. Surfaces connected by Fe atoms are shown to visualize distortions among sub-symmetry transformations better. The C2 (5) symmetry skeleton as shows a great distortion of Fe nets surrounding the Sm center compared with P4/mmm (123) and Immm (71).

The machine learning recommendation system has two central components: (1) the prediction model and (2) the acquisition function. The prediction method used in our framework is the Gaussian process, in which the formation energy is set as the target variable, and the orbital field matrix-based metric learning vector is used as predicting variable.18,41,42 Consequently, structures in the sub-symmetry pool with the lowest predicted formation energy and structures with maximum prediction variance are ranked as the most potential structures used in the first-principle calculation step of the next generation. In detail, structures with the minimum prediction formation energy are expected as the most stable structures–exploitation acquisition function, while structures with the highest predicted variance are expected to enrich the exploration ability of the Gaussian process predictor–exploration acquisition function. Details description and discussion are shown in the previous study in Ref. 18. The genetic algorithm used in this study was inherited from USPEX. For each generation, a structure population consists of 164 crystal structures in which 64 structures are generated by the lattice mutation operator and 100 structures are generated using the machine learning recommender system. By repeatedly performing this process, the population evolved with each element in the population diversified gradually from the original symmetry and, consequently, kept the target of optimizing both the stability and its magnetic properties.

The priority is to maximize the prediction ability of the machine learning estimator using the most limited training data because of the expensive computational resource. Querying new labeled data is equivalent to correcting the form of learned function with respect to target property. As an alternative advantage, following the correction process leads to better insight regarding the phenomena of interest. In this work, a method to localize information on the learned function, monitoring its change to improve the querying data process is introduced.

With the target property as a continuous variable, we consider the learned formation energy function y=f(u), which is called interpretable if it is possible to allocate on the representation space u, where the function meets a predefined condition g. In detail, given a condition g, the probability distribution spanning on the embedding space u is defined as follows:

(1)

with p(u|g) as the probability density at u under g; ui as the location of an observed data point i, i.e., ui=Axi; and h as a tuning kernel width. The indicator 1[] returns 1 if the condition [] is true and 0 otherwise. In the present work, we consider two forms of relevant conditions,

(2)
(3)

where gy(ui) and gxj(ui) intuitively represent a region of interest with potentially stable materials and regions spanned by structures incorporating the maximum value of OFM element xj. Since OFM elements represent information about coordination numbers, the condition gxj(ui) reflect the region of structures owing the corresponding high coordination number. The terminology argmax and argmin are abbreviation of indicating the top 10% highest and lowest value, respectively. Then, we measure the Bhattacharyya coefficient43 between a pair of (gy,gxj) as

(4)

with the integral taken over the space spanned by u. The Bhattacharyya coefficient BC(gy,gxj) measures the probability of joint occurrence between two conditions gy and gxj. Higher BC values indicate a higher possibility to obtain correlation between conditions gy and gxj and vice versa; this makes it easier to understand the meaning of the BC coefficient in identifying overlapping distributions. In the discussion of the results provided in 4.2, we characterize any distribution p(u|g) using a single-level contour representation.

All of the DFT calculations in this study were performed by using the projector augmented wave (PAW) method,44 as implemented in the Vienna ab initio simulation package (VASP),45 within the Perdew–Burke–Ernzerhof (PBE) generalized-gradient approximation (GGA).46 The 4f electrons for all the RE elements are frozen in the core assuming trivalent configuration (Sm_3 pseudopotential. In the first three optimization steps, the lattice constants and atomic positions of all of the study were optimized with a kinetic-energy cutoff for the plane wave expansion 300 eV; the break condition of the electron self-consistent step is set at 1e–4 eV and the smallest allowed spacing between k points is set at 0.5 A1. In the final optimization step, these three values are set at 400 eV, 1e–5 eV and 0.2 A1, respectively. In all steps, the partial occupancies is estimated using Gaussian smearing method with 0.01 eV smearing width.

The formation energy of a given structure s is defined as follows:

(5)

where ΔE[s], E[s], and E[si] are the formation energy, total energy of structure s per formula unit, and simple substance si per atom, respectively. N is the number of atoms in an unit cell. The simple substances were chosen as Im-3m with Fe, R-3m with Sm, Fm-3m with Co with the most converged calculation data from the Open Quantum Materials Database.47,48 A structure whose formation energy lies below or lower than zero, that is, ΔE0, is a potentially formable material in nature, whereas a structure associated with ΔE>0 could be considered unstable. For the competing phases, the stability of the structure should be discussed using the hull distance. In this study, we use the formation energy defined in Eq. (5) as an index for simplicity. The relationship between the experimental material and the hull distance at T=0K has been summarized in Refs. 49 and 50. The stability of the magnets at finite temperature can be found in Ref. 51.

In addition, the total magnetic moment of these materials μ[s] was recalculated because we used an open-core approximation to treat the 4f electrons of Sm, as follows:

(6)

where m[si] is the magnetic moment of atom i, J4fgJ4f[sk] is the correction term with gJ4f as the Lande factor, and J4f is the angular momentum of lanthanide sk. Index i represents all atoms, and index k represents all lanthanide atoms in the structure. The contribution of the 4f electrons of Sm to the magnetization is JgJ=0.714. In this paper, this value is finally converted to magnetization per formula unit, M(T).

In this session, we discuss the structure–stability relationships of SmFe12-based structures extracted using the machine learning-aided genetic algorithm. Table I summarized structures with the minimum formation energy in each corresponding symmetry of SmFe12, SmFe11Co, SmFe11CoB, and SmFe11CoN founded by the system. Overall, our calculated properties show reasonable deviations compared to existing references, e.g., 5% deviation of M, while maintaining the consistent correlations between the substitution–interstitial effects to corresponding physical properties among structure families.

TABLE I.

Minimum formation energy of sub-symmetry structures founded by the machine learning-aided genetic algorithm.

Space groupVolumeFormation energyMagnetizationReference
NameSymmetrynumber(A3)(eV/atom)(T)USPEX idReference
SmFe12 C2/m 12 171.877 0.069 2.115 EA1617 29  
 Fmmm 69 168.128 0.071 1.979 EA1866  
 Immm 71 168.079 0.07 1.978 EA978  
 P4/mmm 123 167.767 0.099 1.820 EA73  
 I4/mmm 139 169.005 0.07 1.989 EA331 31–34  
SmFe11Co Cm 165.678 0.051 1.912 EA1533  
 C2/m 12 169.047 0.045 2.016 EA1539  
 Fmm42 160.151 0.099 1.762 EA18  
 Imm44 166.809 0.049 1.95 EA202 19, 29, 31–36  
SmFe11CoB Cm 173.464 0.024 1.809 EA945 34  
SmFe11CoN Cm 175.783 0.007 1.695 EA449  
 C2/m 12 169.825 −0.054 1.929 EA434  
 Imm44 173.623 −0.058 1.984 EA628 34  
Space groupVolumeFormation energyMagnetizationReference
NameSymmetrynumber(A3)(eV/atom)(T)USPEX idReference
SmFe12 C2/m 12 171.877 0.069 2.115 EA1617 29  
 Fmmm 69 168.128 0.071 1.979 EA1866  
 Immm 71 168.079 0.07 1.978 EA978  
 P4/mmm 123 167.767 0.099 1.820 EA73  
 I4/mmm 139 169.005 0.07 1.989 EA331 31–34  
SmFe11Co Cm 165.678 0.051 1.912 EA1533  
 C2/m 12 169.047 0.045 2.016 EA1539  
 Fmm42 160.151 0.099 1.762 EA18  
 Imm44 166.809 0.049 1.95 EA202 19, 29, 31–36  
SmFe11CoB Cm 173.464 0.024 1.809 EA945 34  
SmFe11CoN Cm 175.783 0.007 1.695 EA449  
 C2/m 12 169.825 −0.054 1.929 EA434  
 Imm44 173.623 −0.058 1.984 EA628 34  

Figure 2 shows the dependence of formation energy on unit-cell volume and magnetization of compounds SmFe12 generated using the proposed machine learning-aided genetic algorithm. This figure shows SmFe12 structures own the formation energy lower than 0.1 (eV/atom) and only space group index larger than 10. There are four notable space group symmetries I/4mmm (139), Immm (71), Fmmm (69), and C2/m (12). Overall, there is a parabolic dependency between the formation energy and the unit-cell volume, as well as the magnetization. The optimal structure owing space group symmetry I/4mmm (139) will get the formation energy 0.07 (eV/atom) at unit-cell volume 168.5 (A3) and magnetization 1.98 (T). Similar trends and the optimal structure appeared with three other symmetry space groups Immm (71), Fmmm (69), and C2/m (12). In this figure, we present the sample skeleton of I/4mmm (139), Immm (71), and C2/m (12). The original I/4mmm (139) appeared with three Fe symmetry sites denoted by 8i, 8j, and 8f Wyckoff index, the SmFe12 structure with Immm (71) symmetry is created by breaking symmetry two Fe 8i sites more than 0.1 (Å)—denoted by white-orange atoms. In the SmFe12 structure with C2/m (12) symmetry, two Fe 8i sites and four Fe 8j sites are broken their symmetry comparing to the original structure.

FIG. 2.

Left: Dependence of formation energy on unit-cell volume and magnetization of compounds SmFe12 generated using the ML-aided genetic algorithm. Right: XRD diffraction of four SmFe12 structures with the optimal formation energy at 0.07 (eV/atom) and the symmetries as I4/mmm (139)—A and B, Immm (71)—C, and C2/m (12)—D.

FIG. 2.

Left: Dependence of formation energy on unit-cell volume and magnetization of compounds SmFe12 generated using the ML-aided genetic algorithm. Right: XRD diffraction of four SmFe12 structures with the optimal formation energy at 0.07 (eV/atom) and the symmetries as I4/mmm (139)—A and B, Immm (71)—C, and C2/m (12)—D.

Close modal

In Fig. 2, it should be noted that there are some atom arrangements in the unit cell of I/4mmm (139) structure, e.g., structure A sharing a similar formation energy achieve 0.07 (eV/atom) but expanding the unit-cell volume to 174.3 (Å3)– 3.4% larger than the optimal formation energy structure with the unit-cell volume at 168.5 (Å3), structure B. With 2.13 (T), the magnetization of the structure A shows 7.5% higher than the B’s magnetization.

We pickup four SmFe12 structures with the formation energy at 0.07 (eV/atom) denoted by A, B, C, and D for studying the difference of x-ray diffraction patterns with Cu-Kα, λ=1.5406 (Å) radiation. Two structures, A and B, share a similar symmetry I4/mmm (139), but the magnetization as M=2.132 (T) and 1.989 (T), respectively. Two other structures, C and D, are associated with the Immm (71) and C2/m (12) symmetries, respectively. One might observe that x-ray diffraction of structure A appears with an approximately single peak in range 2θ from 42° to 44°, while three other structures appear with two separated peaks. Therefore, a slight distortion of the original SmFe12 may also cause the minor peak appearing in 44.5°, which is similar to the peak of αFe in XRD spectra of synthesized structures as previous reports.52 Also, the relative intensity of two peaks at 37 and 38 of structure A shows more consistent with report in Refs. 53 and 20 than the other.

Figure 3 shows the dependence of formation energy on unit-cell volume and magnetization of SmFe11Co structures. By limiting the space group symmetry of founded structures, there are three separate correlations of the formation energy on the unit-cell volume for three substituted Co atom positions. Overall, all substituted structures are more formable than the original SmFe12 structure, even the original space group symmetry I/4mmm (139) broken down to C2/m (12) and Imm2 (44). For a given value of unit-cell volume, the SmFe11Co structure with the Co atom occupied Fe 8f site always shows the lowest formation energy, while the structure with Co atom substituted to Fe 8i always has the highest formation energy. However, while there is no difference between magnetization values for structures with Co atom substituted to Fe 8i and 8j sites, the Co atom substitute to 8f site can achieve significantly larger magnetization.18 Regarding the results of magnetization dependence on substituted elements, almost all substituted structures acquired lower magnetization values than the original SmFe12 structure, except for the substitution of Co at the 8f position of the Fe site. This result is consistent with other first-principle calculation code Quantum Materials Simulator (QMAS);31,35 OpenMX19 or synthesis reports of Sm(Fe0.9Co0.1)12 and Sm(Fe0.8Co0.2)12 in Refs. 32 and 36. From experimental results, authors in Ref. 32 reported that Sm(Fe0.8Co0.2)12 had not yet been stabilized in the bulk form, which is consistent with the positive formation energy appearing with Co substitution from our calculation, although they had lower energy than the host structures.

FIG. 3.

Dependence of formation energy on unit-cell volume and magnetization of compounds SmFe11Co and SmFe12 structures generated using the machine learning-aided genetic algorithm.

FIG. 3.

Dependence of formation energy on unit-cell volume and magnetization of compounds SmFe11Co and SmFe12 structures generated using the machine learning-aided genetic algorithm.

Close modal

In Fig. 4, we show the results of SmFe11CoB and SmFe11CoN in comparison with the original formation energy and magnetization of the SmFe12 structure with I/4mmm (139) symmetry. While the B interstitial method— (gray shade) may create the structure with lower formation energy—0.025 (eV/atom), this method also decreases the optimal magnetization value from 2.0 (T) of the SmFe11Co to 1.8 (T). On the other hand, a part of N interstitial structures with Cm (8) symmetry shows remaining formation energy at 0.015 (eV/atom) for a wide range of magnetization from 1.5 to 1.81 (T). However, the other part of N interstitial structures with Imm2 (44) and C2/m (12) symmetries can achieve negative formation energy up to 0.06 (eV/atom) but keep the optimal magnetization at a high value from 1.85 to 1.98 (T). In the lower panel of Fig. 4, we show formation energy and magnetization of SmFe11CoX structures with X as Al, Ti, Cu, Co, Zn, Mo, and Ga collected from Ref. 18 as reference. All SmFe10CoX structures appear on the left region of the graph—the region with low magnetization compared with the original SmFe12 structure. To summarize, SmFe11CoN structures with Imm2 (44) and C2/m (12) symmetries surpass all other structures in the ability to maintain low formation energy and high magnetization.

FIG. 4.

Upper: Formation energy and magnetization of compounds SmFe12, SmFe11Co, SmFe11CoB, and SmFe11CoN generated using the machine learning-aided genetic algorithm. Lower: Comparison with available data of SmFe11CoX2 and SmFe11XCo2 structures with X as Al, Ti, Cu, Co, Zn Mo, and Ga.18 

FIG. 4.

Upper: Formation energy and magnetization of compounds SmFe12, SmFe11Co, SmFe11CoB, and SmFe11CoN generated using the machine learning-aided genetic algorithm. Lower: Comparison with available data of SmFe11CoX2 and SmFe11XCo2 structures with X as Al, Ti, Cu, Co, Zn Mo, and Ga.18 

Close modal

In this part, we show results in interpreting the structure–stability relationship of the most prominent SmFe11CoN structure family using the embedding space by metric learning for the kernel regression (MLKR) method. Figure 5 shows the space of SmFe11CoN structures learned by embedding information on the formation energy. Appropriated kernel density estimation parameters are applied for smoothing interpolation in the embedding space. As the nature of the metric learning method, the formation energy shown in the background in Fig. 5(b) smoothly transverse through calculated structures in which neighbor structures share both similar geometrical information and formation energy. In Fig. 5(c), we show a similar space, but the background color is filled by the calculated magnetization M.

FIG. 5.

(a) Bhattacharyya coefficient (BC) score measuring overlapping between regions of SmFe11CoN structures with high stability and region of structures with maximum coordination number measured using the orbital field matrix. The BC score is shown in every generation of the USPEX program. Metric learned embedding space visualizes SmFe11CoN structures with background filled by first-principle calculated (b) formation energy and (c) magnetization. (d) SmFe11CoN structures with different positions of occupied cobalt substitution sites and nitrogen interstitial sites.

FIG. 5.

(a) Bhattacharyya coefficient (BC) score measuring overlapping between regions of SmFe11CoN structures with high stability and region of structures with maximum coordination number measured using the orbital field matrix. The BC score is shown in every generation of the USPEX program. Metric learned embedding space visualizes SmFe11CoN structures with background filled by first-principle calculated (b) formation energy and (c) magnetization. (d) SmFe11CoN structures with different positions of occupied cobalt substitution sites and nitrogen interstitial sites.

Close modal

In the embedding space, we represent the condition of Bhattacharyya measure with the red contour indicating the region of minimized formation energy—Fig. 5(b), and maximized the magnetization—Fig. 5(c). Other contours indicating regions of structures with the maximum of the corresponding coordination number. For example, (f6,p3) contour indicates regions of structures owning with the maximum value of coordination number of N atom—p3, surrounding Sm atom—f6. In Fig. 5(b), there are three separated regions with minimum formation energy that represent for three different geometrical arrangements of substituted Co site and N interstitial site that cause similar stability level.

In Fig. 5(a), we show the Bhattacharyya coefficient score measuring overlapping between regions of SmFe11CoN structures with the maximal OFM condition and the region of minimum formation energy. As the most appearance in this metric, highly stable SmFe11CoN structures are associated with a high coordination number of Fe/Co surrounding the Fe site. Also, there is a higher possibility of stability for structures with an N atom surrounding the Sm center rather than the Co or Fe atom, i.e., the Bhattacharyya coefficient for (f6,p3) significantly large than (d7,p3) and (d6,p3), respectively. Apparently shown in the embedding map in Fig. 5(b), the (f6,p3) contour show higher overlapping to stable structure region compared to other contours. The region allocated by (f6,p3) also show the maximal calculated magnetization.

Figure 5(d) shows four reference structures, EA1526, EA1593, EA628, and EA1402, as the representative structures for geometrical information in the three most stable regions in Fig. 5(b). The region of maximal (f6,p3) represented by EA628 and EA1402 structures own relatively higher magnetization—larger than 1.8 (T) compared to two other stable regions. It is the region with structures showing N atoms placed in the middle of two Sm atoms (2p), compared with the other regions where the N atom interstate among other Fe and Co atoms. It is exclusively the region with symmetry group found by Imm2 (44) with Co site substituted in 8i position—EA628 structure, and C2/m (12) with Co site substituted in 8f position—EA1402 structure. The preferred position of the N interstitial site in the center of the Fe and Sm octahedron inducing strong magnetocrystalline anisotropy on Sm is consistent with previous reports.34,54 Further investigation of the geometrical shape of other stable structures suggests that interstitial positions of N rather than the middle of two Sm atoms cease more considerable distortion to the original SmFe12 skeleton with the I4/mmm (139) space group symmetry.

In Fig. 6, we summarize the correlation between formation energy, magnetization, and the density of three structure families investigated SmFe12, SmFe11Co, and SmFe11CoN. Regardless of the elemental constituent of each family, there is a general trend that by reducing the atomic density of the structure, the corresponding magnetization linearly increases while the formation energy decrease until a stable position.

FIG. 6.

Dependence of formation energy (left axis) and magnetization (right axis) on atomic density of SmFe12, SmFe11Co, and SmFe11CoN.

FIG. 6.

Dependence of formation energy (left axis) and magnetization (right axis) on atomic density of SmFe12, SmFe11Co, and SmFe11CoN.

Close modal

In this work, we introduce a machine learning-aided genetic algorithm structure generation method to investigate the correlation between geometrical information, stability, and magnetization of SmFe12-based structures. For every population of genetic algorithm generation, a pool of structures is created using the sub-symmetry perturbation method. The gaussian process as predictor taking the embedded orbital field matrix representation as structure fingerprint has been applied to ranking the most potential stability structures for first-principle calculation. As a result, the original structure SmFe12, well-known with tetragonal I4/mmm skeleton structures, is investigated with distortions of the unit-cell lattice parameter and individual sites. An anomaly atom arrangement of the I4/mmm SmFe12 structure is found out of the parabolic dependence between the formation energy and magnetization. Structures with nitrogen interstitial position in the middle of two Samarium sites show outperform all other structures in both ability of stabilization and remaining high magnetization order of the original structure. Lastly, metric learning embedding space brings fruitful insight into the correlation between this structure family’s geometrical arrangement, stability, and magnetization.

This work was supported by the Ministry of Education, Culture, Sports, Science, and Technology of Japan (MEXT) with the Program for Promoting Research on the Supercomputer Fugaku (DPMSD), JSPS KAKENHI Grants 20K05301 and JP19H05815 (Grants-in-Aid for Scientific Research on Innovative Areas Interface Ionics), 21K14396 (Grant-in-Aid for Early- Career Scientists), and 20K05068, Japan.

The authors have no conflicts to disclose. Hereby, I, D.-N.N consciously assure that the manuscript “Machine learning-aided genetic algorithm in investigating the structure–property relationship of SmFe12-based structures” is the authors’ own original work, which has not been previously published elsewhere. The paper reflects the author’s own research and analysis in a truthful and complete manner. Besides, the paper properly credits the meaningful contributions of co-authors and co-researchers and is appropriately placed in the context of prior and existing research.

Informed consent was obtained from all individual participants included in the study. The participant has consented to the submission of the case report to the journal.

Duong-Nguyen Nguyen: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Funding acquisition (lead); Investigation (lead); Methodology (equal); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (lead). Hieu-Chi Dam: Conceptualization (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal).

Dataset of SmFe12, SmFe11X with X = {Co, CoB, CoN} including vasp calculations, OFM descriptor is openly available in Zenodo at https://doi.org/10.5281/zenodo.7318435, Ref. 55. Dataset of SmFe12αβXαYβ structures with X,Y as Mo, Zn, Co, Cu, Ti, Al, Ga, and α+β<4 including vasp calculations, OFM descriptor is openly available in Zenodo at https://doi.org/10.5281/zenodo.5763325, Ref. 56.

1.
D. S.
Ranasinghe
,
A.
Perera
, and
R. J.
Bartlett
, “
A note on the accuracy of KS-DFT densities
,”
J. Chem. Phys.
147
,
204103
(
2017
).
2.
M. G.
Medvedev
,
I. S.
Bushmarinov
,
J.
Sun
,
J. P.
Perdew
, and
K. A.
Lyssenko
, “
Density functional theory is straying from the path toward the exact functional
,”
Science
355
,
49
52
(
2017
).
3.
S.
Hammes-Schiffer
, “
A conundrum for density functional theory
,”
Science
355
,
28
29
(
2017
).
4.
G.
Graziano
, “
Quantum chemistry: DFT’s midlife crisis
,”
Nat. Rev. Chem.
1
,
0019
(
2017
).
5.
A. D.
Becke
, “
Perspective: Fifty years of density-functional theory in chemical physics
,”
J. Chem. Phys.
140
,
18A301
(
2014
).
6.
R. O.
Jones
, “
Density functional theory: Its origins, rise to prominence, and future
,”
Rev. Mod. Phys.
87
,
897
923
(
2015
).
7.
A.
Gavezzotti
, “
Are crystal structures predictable?
,”
Acc. Chem. Res.
27, 10
,
309
314
(
1994
).
8.
Y. K.
Takahashi
,
H.
Sepehri-Amin
, and
T.
Ohkubo
, “
Recent advances in SmFe12-based permanent magnets
,”
Sci. Technol. Adv. Mater.
22
,
449
460
(
2021
).
9.
P.
Tozman
,
Y.
Takahashi
,
H.
Sepehri-Amin
,
D.
Ogawa
,
S.
Hirosawa
, and
K.
Hono
, “
The effect of Zr substitution on saturation magnetization in (Sm1xZrx)(Fe0.8Co0.2)12 compound with the ThMn12 structure
,”
Acta Mater.
178
,
114
121
(
2019
).
10.
S.
Suzuki
,
T.
Kuno
,
K.
Urushibata
,
K.
Kobayashi
,
N.
Sakuma
,
K.
Washio
,
M.
Yano
,
A.
Kato
, and
A.
Manabe
, “
A new magnet material with ThMn12 structure: (Nd1xZrx)(Fe1yCoy)11+zTi1zNα (α=0.61.3)
,”
J. Magn. Magn. Mater.
401
,
259
268
(
2016
).
11.
R.
Coehoorn
, “
Electronic structure and magnetism of transition-metal-stabilized YFE12xMx intermetallic compounds
,”
Phys. Rev. B
41
,
11790
11797
(
1990
).
12.
K.
Buschow
, “
Permanent magnet materials based on tetragonal rare earth compounds of the type RFe12xMx
,”
J. Magn. Magn. Mater.
100
,
79
89
(
1991
).
13.
Y.
Wang
and
G.
Hadjipanayis
, “
Magnetic properties of Sm-Fe-Ti-V alloys
,”
J. Magn. Magn. Mater.
87
,
375
378
(
1990
).
14.
T.
Fukazawa
,
H.
Akai
,
Y.
Harashima
, and
T.
Miyake
, “
First-principles study of spin-wave dispersion in Sm(Fe1xCox)12
,”
J. Magn. Magn. Mater.
469
,
296
301
(
2019
).
15.
T.
Fukazawa
,
Y.
Harashima
,
Z.
Hou
, and
T.
Miyake
, “
Bayesian optimization of chemical composition: A comprehensive framework and its application to RFe12-type magnet compounds
,”
Phys. Rev. Mater.
3
,
053807
(
2019
).
16.
A.
Schönhöbel
,
R.
Madugundo
,
O. Y.
Vekilova
,
O.
Eriksson
,
H.
Herper
,
J.
Barandiarán
, and
G.
Hadjipanayis
, “
Intrinsic magnetic properties of SmFe12xVx alloys with reduced V-concentration
,”
J. Alloys Compd.
786
,
969
974
(
2019
).
17.
T.
Miyake
,
Y.
Harashima
,
T.
Fukazawa
, and
H.
Akai
, “
Understanding and optimization of hard magnetic compounds from first principles
,”
Sci. Technol. Adv. Mater.
22
,
543
556
(
2021
), pMID: 34552388.
18.
D.-N.
Nguyen
,
H.
Kino
,
T.
Miyake
, and
H.-C.
Dam
, “Explainable active learning in investigating structure-stability of SmFe12-α-βXαYβ structures X, Y {Mo, Zn, Co, Cu, Ti, Al, Ga},”
MRS Bulletin
(published online, 2022).
19.
M.
Matsumoto
,
T.
Hawai
, and
K.
Ono
, “
(Sm,Zr)Fe12xMx (M=Zr,Ti,Co) for permanent-magnet applications: Ab initio material design integrated with experimental characterization
,”
Phys. Rev. Appl.
13
,
064028
(
2020
).
20.
A.
Gabay
and
G.
Hadjipanayis
, “
Recent developments in RFe12-type compounds for permanent magnets
,”
Scr. Mater.
154
,
284
288
(
2018
).
21.
M.
Hagiwara
,
N.
Sanada
, and
S.
Sakurada
, “
Effect of Y substitution on the structural and magnetic properties of Sm(Fe0.8Co0.2)11.4Ti0.6
,”
J. Magn. Magn. Mater.
465
,
554
558
(
2018
).
22.
A. M.
Gabay
,
A.
Martín-Cid
,
J. M.
Barandiaran
,
D.
Salazar
, and
G. C.
Hadjipanayis
, “
Low-cost Ce1xSmx(Fe, Co, Ti)12 alloys for permanent magnets
,”
AIP Adv.
6
,
056015
(
2016
).
23.
A.
Gabay
and
G.
Hadjipanayis
, “
Mechanochemical synthesis of magnetically hard anisotropic RFe10Si2 powders with R representing combinations of Sm, Ce and Zr
,”
J. Magn. Magn. Mater.
422
,
43
48
(
2017
).
24.
N.
Sakuma
,
S.
Suzuki
,
T.
Kuno
,
K.
Urushibata
,
K.
Kobayashi
,
M.
Yano
,
A.
Kato
, and
A.
Manabe
, “
Influence of Zr substitution on the stabilization of ThMn12-type (Nd1αZrα)(Fe0.75Co0.25)11.25Ti0.75N1.21.4 (α=00.3) compounds
,”
AIP Adv.
6
,
056023
(
2016
).
25.
Y.
Horiuchi
,
M.
Hagiwara
,
M.
Endo
,
N.
Sanada
, and
S.
Sakurada
, “
Influence of intermediate-heat treatment on the structure and magnetic properties of iron-rich Sm(CoFeCuZr)z sintered magnets
,”
J. Appl. Phys.
117
,
17C704
(
2015
).
26.
Y.
Hirayama
,
A.
Panda
,
T.
Ohkubo
, and
K.
Hono
, “
High coercivity Sm2Fe17N3 submicron size powder prepared by polymerized-complex and reduction–diffusion process
,”
Scr. Mater.
120
,
27
30
(
2016
).
27.
Y.
Guo
,
J.
You
,
W.
Pei
,
Y.
Qu
,
D.
Zhao
, and
Z.
Yang
, “
Effect of (C2H5)3NBH3 content on microstructure and properties of Nd-Fe-B nanoparticles prepared by chemical and reduction-diffusion method
,”
J. Alloys Compd.
777
,
850
859
(
2019
).
28.
K.
Takagi
,
Y.
Hirayama
,
S.
Okada
,
A.
Hosokawa
, and
W.
Yamaguchi
, “
Recent research trend in powder process technology for high-performance rare-earth permanent magnets
,”
KONA Powder Part J.
advpub
,
2023007
(
2022
).
29.
T.
Ishikawa
,
T.
Fukazawa
, and
T.
Miyake
, “
Monoclinic YFe12 phases predicted from first principles
,”
Phys. Rev. Mater.
4
,
104408
(
2020
).
30.
D.
Wang
,
S.
Liou
,
P.
He
,
D.
Sellmyer
,
G.
Hadjipanayis
, and
Y.
Zhang
, “
SmFe12 and SmFe12Nx films fabricated by sputtering
,”
J. Magn. Magn. Mater.
124
,
62
68
(
1993
).
31.
Y.
Harashima
,
K.
Terakura
,
H.
Kino
,
S.
Ishibashi
, and
T.
Miyake
, “First-principles study of structural and magnetic properties of R(Fe,Ti)12 and R(Fe,Ti)12N (R = Nd, Sm, Y),” in Proceedings of Computational Science Workshop 2014 (CSW2014) (Physical Society of Japan, 2014).
32.
Y.
Hirayama
,
Y.
Takahashi
,
S.
Hirosawa
, and
K.
Hono
, “
Intrinsic hard magnetic properties of Sm(Fe1xCox)12 compound with the ThMn12 structure
,”
Scr. Mater.
138
,
62
65
(
2017
).
33.
D.
Ogawa
,
T.
Yoshioka
,
X.
Xu
,
Y.
Takahashi
,
H.
Tsuchiura
,
T.
Ohkubo
,
S.
Hirosawa
, and
K.
Hono
, “
Magnetic anisotropy constants of ThMn12-type Sm(Fe1xCox)12 compounds and their temperature dependence
,”
J. Magn. Magn. Mater.
497
,
165965
(
2020
).
34.
D.
Odkhuu
,
T.
Ochirkhuyag
, and
S. C.
Hong
, “
Enhancing energy product and thermal stability of SmFe12 by interstitial doping
,”
Phys. Rev. Appl.
13
,
054076
(
2020
).
35.
I.
Dirba
,
Y.
Harashima
,
H.
Sepehri-Amin
,
T.
Ohkubo
,
T.
Miyake
,
S.
Hirosawa
, and
K.
Hono
, “
Thermal decomposition of ThMn12-type phase and its optimum stabilizing elements in SmFe12-based alloys
,”
J. Alloys Compd.
813
,
152224
(
2020
).
36.
S.
Hirosawa
, “
Permanent magnets beyond Nd-Dy-Fe-B
,”
IEEE Trans. Magn.
55
,
1
6
(
2019
).
37.
S.
Fredericks
,
K.
Parrish
,
D.
Sayre
, and
Q.
Zhu
, “
Pyxtal: A python library for crystal structure generation and symmetry analysis
,”
Comput. Phys. Commun.
261
,
107810
(
2021
).
38.
M. I.
Aroyo
,
J. M.
Perez-Mato
,
C.
Capillas
,
E.
Kroumova
,
S.
Ivantchev
,
G.
Madariaga
,
A.
Kirov
, and
H.
Wondratschek
, “
Bilbao crystallographic server: I. Databases and crystallographic computing programs
,”
Z. Kristallog. Cryst. Mater.
221
,
15
27
(
2006
).
39.
M. I.
Aroyo
,
J.
Manuel Perez-Mato
,
C.
Capillas
, and
H.
Wondratschek
, “
The Bilbao Crystallographic Server
,”
International Tables for Crystallography
(Wiley, 2011), Vol. A1, Chap. 1.7, pp. 57–69.
40.
M. I.
Aroyo
,
A.
Kirov
,
C.
Capillas
,
J. M.
Perez-Mato
, and
H.
Wondratschek
, “
Bilbao crystallographic server. II. Representations of crystallographic point groups and space groups
,”
Acta Crystallogr. Sect. A
62
,
115
128
(
2006
).
41.
T.
Lam Pham
,
H.
Kino
,
K.
Terakura
,
T.
Miyake
,
K.
Tsuda
,
I.
Takigawa
, and
H.
Chi Dam
, “
Machine learning reveals orbital interaction in materials
,”
Sci. Technol. Adv. Mater.
18
,
756
765
(
2017
).
42.
T.-L.
Pham
,
N.-D.
Nguyen
,
V.-D.
Nguyen
,
H.
Kino
,
T.
Miyake
, and
H.-C.
Dam
, “
Learning structure-property relationship in crystalline materials: A study of lanthanide–transition metal alloys
,”
J. Chem. Phys.
148
,
204106
(
2018
).
43.
A. K.
Bhattacharyya
, “
On a measure of divergence between two statistical populations defined by their probability distribution
,”
Bull. Calcutta Math. Soc.
35
,
99
110
(
1943
).
44.
G.
Kresse
and
D.
Joubert
, “
From ultrasoft pseudopotentials to the projector augmented-wave method
,”
Phys. Rev. B
59
,
1758
1775
(
1999
).
45.
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
11186
(
1996
).
46.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
3868
(
1996
).
47.
J. E.
Saal
,
S.
Kirklin
,
M.
Aykol
,
B.
Meredig
, and
C.
Wolverton
, “
Materials design and discovery with high-throughput density functional theory: The open quantum materials database (OQMD)
,”
J. Mater.
65
,
1501
1509
(
2013
).
48.
S.
Kirklin
,
J. E.
Saal
,
B.
Meredig
,
A.
Thompson
,
J. W.
Doak
,
M.
Aykol
,
S.
Rühl
, and
C.
Wolverton
, “
The open quantum materials database (OQMD): Assessing the accuracy of dft formation energies
,”
npj Comput. Mater.
1
,
15010
(
2015
).
49.
W.
Sun
,
S. T.
Dacek
,
S. P.
Ong
,
G.
Hautier
,
A.
Jain
,
W. D.
Richards
,
A. C.
Gamst
,
K. A.
Persson
, and
G.
Ceder
, “
The thermodynamic scale of inorganic crystalline metastability
,”
Sci. Adv.
2
,
e1600225
(
2016
).
50.
Y.
Wu
,
P.
Lazic
,
G.
Hautier
,
K.
Persson
, and
G.
Ceder
, “
First principles high throughput screening of oxynitrides for water-splitting photocatalysts
,”
Energy Environ. Sci.
6
,
157
168
(
2013
).
51.
G.
Xing
,
T.
Ishikawa
,
Y.
Miura
,
T.
Miyake
, and
T.
Tadano
, “Lattice dynamics effects on finite-temperature stability of R1xFex (R = Y, Ce, Nd, Sm, and Dy) alloys from first principles,” arxiv (2021), https://arxiv.org/abs/2102.02590.
52.
H.-D.
Qian
,
J. T.
Lim
,
J.-W.
Kim
,
Y.
Yang
,
K. M.
Cho
,
J.
Park
, and
C.-J.
Choi
, “
Phase transformation and magnetic properties of fully dense Sm(Fe0.8Co0.2)11Ti bulk magnets
,”
Scr. Mater.
193
,
17
21
(
2021
).
53.
G.
Hadjipanayis
,
A.
Gabay
,
A.
Schönhöbel
,
A.
Martín-Cid
,
J.
Barandiaran
, and
D.
Niarchos
, “
ThMn12-type alloys for permanent magnets
,”
Engineering
6
,
141
147
(
2020
).
54.
M.
Akayama
,
H.
Fujii
,
K.
Yamamoto
, and
K.
Tatami
, “
Physical properties of nitrogenated RFe11Ti intermetallic compounds (R = Ce, Pr and Nd) with ThMn12-type structure
,”
J. Magn. Magn. Mater.
130
,
99
107
(
1994
).
55.
N. D.
Nguyen
and
D. H.
Chi.
(
2022
).
SmFe12, SmFe11Co, Smfe11CoB, Smfe11CoN structures OFM representation, magnetization and formation energy [Data set]
. Zenodo.
56.
N. D.
Nguyen
and
D. H.u
Chi.
(
2021
).
SmFe12 substituted structures with OFM representation, magnetization and formation energy
.