Site preferences of alloying transition metal elements in Ni-based superalloy: A first-principles study
Lu Baokun1, Wang Chong-Yu1, †, Du Zhihui2
Department of Physics, Tsinghua University, Beijing 100084, China
Department of Computer Science and Technology, Tsinghua University, Beijing 100084, China

 

† Corresponding author. E-mail: cywang@mail.tsinghua.edu.cn

Project supported by the National Key R&D Program of China (Grant Nos. 2017YFB0701501, 2017YFB0701502, and 2017YFB0701503).

Abstract

Atomistic characterization of chemical element distribution is crucial to understanding the role of alloying elements for strengthening mechanism of superalloy. In the present work, the site preferences of two alloying elements XY in γ-Ni of Ni-based superalloy are systematically studied using first-principles calculations with and without spin-polarization. The doping elements X and Y are chosen from the 27 kinds of 3d, 4d, 5d group transition metals (Sc, Ti, V, Cr, Mn, Fe, Co, Cu, Zn, Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, Cd, Hf, Ta, W, Re, Os, Ir, Pt, Au) and Al. We find that the spin-polarized calculations for Re–Re, Re–Ru, Re–Cr, Ru–Cr show a strong chemical binding affinity between the solute elements and are more consistent with the experimental results. The binding energies of pairs between the 28 elements have an obvious periodicity and are closely related the electronic configuration of the elements. When the d-electrons of the element are close to the half full-shell state, two alloying elements possess attractive binding energies, reflecting the effect of the Hund’s rule. The combinations of early transition metals (Sc, Ti, V, Y, Zr, Nb, Hf, Ta) have a repulsive interaction in γ-Ni. These results offer insights into the role of alloying elements for strengthening mechanism of superalloy.

1. Introduction

Ni-base superalloys are unique high temperature materials, which display excellent high-temperature properties, such as strength, ductility, fracture toughness, fatigue resistance, creep resistance, and oxidation resistance.[1,2] Increasing demands for applications of such materials, mostly used in turbine blades of aircraft jet engines and land-based power generators, drive continued development of these alloys to perform in extreme conditions. The superalloys are characterized by cubical γ′ precipitates (L12, ordered face-centered cubic (fcc), intermetallic compounds Ni3Al) embedded in the γ matrix (disordered fcc, solid solution Ni).[1] In order to strengthen the single-crystal superalloys, nearly 10 elements are usually doped to improve the high temperature mechanical properties and their lifetime. Transition alloying elements Ti, Co, Cr, Nb, Mo, Ru, Ta, Hf, W, and Re are added to enhance the high-temperature properties. The elements Al and Ta are enriched in the γ′ phase, while the elements Co, Cr, Mo, W, Re, and Ru are enriched in the γ matrix phase. Solid solution strengthening of the γ matrix is one key factor for improving the creep strength of Ni-base superalloys at high temperatures. Creep processes are largely controlled by the thermally activated motion of dislocations in the matrix phase, which requires dislocation climb at the γ/γ′ interfaces. Elemental diffusion strongly influences the dislocation motion, thus it is important to consider the interaction of the solute elements. Among these alloying elements, Re remarkably improves the strength and creep resistance. The content of Re is considered as one of the characteristics in different-generation single-crystal superalloys, and in the newly developed fourth- and fifth-generation single-crystal superalloys.

However, no widely convinced mechanism has been conferred for the strengthening effect of Re until now. It was proposed that Re is present in the form of clusters in γ and acts as efficient obstacles for dislocation motion. Based on the results of one-dimensional atom probe (1DAP), Re clustering postulate was first suggested by Blavette et al.[35] Wanderka et al.[6] further used atom probe tomography (APT) to study the superalloy CMSX-4 and also concluded the formation of Re cluster. Rüsing et al.[7] used the more advanced three-dimensional atom probe (3DAP) and concluded that Re forms clusters with sizes of about 1 nm and mutual average distances of 20 nm. Mottura et al.[8] drew a contrary conclusion by using APT method, which took into account the detection efficiency and positional scatter of the atoms, and a purpose-built algorithm was used to analyze the atom probe datasets. They showed that the fluctuation in the ladder diagrams may be due to the random variation in the solute atoms and no solute clusters were detected in the Ni–Re binary alloys or the CMSX-4 single-crystal superalloys.[8,9] The experimental x-ray absorption fine structure (EXAFS) spectra were also compared to the simulated spectra, and it was shown that the experimental Ni–Re alloy does not display the double peaks in the spectra and Re will not form clusters.[10]

In addition to these experimental studies, simulation methods were also used to check whether Re forms clusters or not in superalloys. The ab initio density functional theory simulations showed strong negative binding energies between Re–Re nearest neighbor pairs, and the repulsion energy for Re–Re pairs is rapidly reduced as the atoms are placed further apart.[10,11] Zhu et al.[12] used molecular dynamics (MD) simulations to calculate the binding energies of Re clusters in the Ni–Al–Re system and the results showed strong binding energies between Re atoms and indicated that Re tends to form clusters in the γ phase.

From a physical point of view, the co-segregation of Re with other alloying elements may be responsible for the strengthening effect of Re. The clusters built up by a couple of Re-solute atoms may play a major role in improving the creep strength of Re-containing superalloys.[13] Though not much information is available with the synergistic effects of Re with other alloying elements, some studies do give some indications of the synergy. Smith[14] studied a series of simple model alloys and showed that Cr strongly affects the partitioning of Re. And Re could not strengthen the superalloys without the addition of Cr. The following work of Chen et al.[15] showed similar behavior of the effect of Cr on Re and some bright contrast regions in the γ phase near the γ/γ′ interface were found. Recently, Ge et al.[16] studied the distribution of Re in a second-generation single-crystal superalloy DD6 using high-angle annular dark-field scanning transmission electron microscope (HAADF–STEM) with the energy-dispersive spectroscopy (EDS) mapping. They suggested that Re and W may form clusters with sizes of about 1–2 nm in the γ phase close to the γ/γ′ interface in crept superalloys, while no such clusters could be found near the γ/γ′ interface before creep test. Co-segregation of Re with Co and Cr was observed at the tips of protrusions at the γ/γ′ interface. As a result of coupling between the interfacial dislocation motion and the co-segregation of Re with Co and Cr, the creep rate in the steady creep regime was slowed down.[17] Yu et al.[18] used transmission electron microscopy and first-principles calculations to show that Re and Ru have a strong interaction, which results in the repartitioning of Re to the γ phase. Huang et al.[19] demonstrated that correlated element–element binding energies play a key role in the observed Ni retention-excesses at the γ-Ni/γ′-Ni3Al interfaces during aging and evolution of the precipitates. In order to obtain deeper insight into the electronic mechanism of the element distribution, further theoretical calculations are needed to understand the interaction of Re element with other alloying elements at atomic-scale.

In the present work, via studying the binding energy of two doping elements, the site preferences of two alloying elements XY in γ-Ni of Ni-based superalloy are systematically studied using first-principles calculations with and without spin-polarization. The doping elements X and Y are chosen from the 27 kinds of 3d, 4d, 5d group transition metals (Sc, Ti, V, Cr, Mn, Fe, Co, Cu, Zn, Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, Cd, Hf, Ta, W, Re, Os, Ir, Pt, Au) and Al. We find that the spin-polarized calculations for Re–Re, Re–Ru, Re–Cr, Ru–Cr show a strong chemical binding affinity between the solute elements and more consistent with the experimental results. The binding energies of pairs between the 28 elements have an obvious periodicity and are closely related the electronic configurations of the elements. When d-electrons of an element is close to the half full-shell state, the doping element possesses an attractive binding energy, reflecting the effect of the Hund’s rule. Meanwhile, the combination of early transition metals (Sc, Ti, V, Y, Zr, Nb, Hf, Ta) has a repulsive interaction in γ-Ni.

2. Methodology
2.1. Computational method

The first-principles calculations are based on electronic density-functional theory (DFT) and have been carried out using Vienna ab initio simulation package (VASP).[2023] The electronic wave functions are expanded in plane waves with a kinetic-energy cutoff of 350 eV. For the DFT calculations, the projector augmented wave method[24] and the generalized gradient approximation of Perdew, Burke, and Ernzerhof[25] are used. A combination of conjugate gradient and quasi-Newton optimization methods is employed in this work to achieve faster geometry optimization due to high computational cost for calculating these large supercell systems. The volumes of all structures and the atomic arrangements are fully relaxed. The convergence criteria for the total energy and forces are chosen as 10−5 eV and 0.02 eV/Å, respectively. The Brillouin zone integrations are performed using Monkhorst–Pack[26]k-point meshes, a 5 × 5 × 5 k-point mesh is used for the 3 × 3 × 3 supercells, which is found to be sufficient to give fully converged results. We use the first-order Methfessel–Paxton[27] technique with the smearing parameter of 0.1 eV. The set of criteria achieves the generalized sufficient convergence condition. Spin-polarized and non-spin-polarized calculations are performed in the present study to take into account the effect of magnetism.

2.2. Computational model

Throughout this work, a 108 atom supercell model is used as shown in Fig. 1. There are 108 possible positions to replace an atom, and all positions are equivalent based on the crystal symmetry. When element X is doped at the corner site (0,0,0), there are a total of 107 possible sites for the second Y element to occupy. Taking into account the symmetry factors, there are a total of 9 non-equivalent models. The difference between the 9 non-equivalent models is the distance between the two replacement atoms. The distance between the two doping elements at an ith nearest neighbor (NN), the symmetry of the systems, the doping position of the second element Y, the number of equivalent models, and its weight are shown in Table 1. Doping elements X and Y are chosen from the 27 kinds of 3d, 4d, and 5d group transition metals (Sc, Ti, V, Cr, Mn, Fe, Co, Cu, Zn, Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, Cd, Hf, Ta, W, Re, Os, Ir, Pt, Au) and Al. X = Ni and Y = Ni correspond to the Ni107X and Ni107Y binary systems, respectively. There are a total of 784 possibilities, Ni106XY and Ni106YX are equivalent systems, so the number of non-equivalent systems is 406, including 378 ternary systems (Ni106XY) and 28 binary systems (Ni106XX).

Fig. 1. (color online) The nine nearest neighbor pairs considered in 3 × 3 × 3 108 atoms supercell model.
Table 1.

The nine nearest neighbor pairs are considered in 108 atoms model.

.
2.3. Binding energy

The binding energy between two doping defects at an ith NN distance is calculated by subtracting the total energy of the system with the two doping defects at the ith NN distance from the total energy of the system with the two doping defects at infinity.[19] It is given by

In a 3 × 3 × 3 supercell system, the 11th NN distance is far enough and the binding energy of the solute pairs is expected to reach 0 eV. The binding energy can be expressed as
Also, if the supercell is large enough, the binding energy can be expressed as[11]
In Eqs. (1) and (3), E(Nin−1X) and E(Nin−1Y) are the total energies of the supercell with a single X or Y atom, EithNN(Nin−2XY) is the total energy of the supercell with two doping atoms at an ith NN distance, E(Nin) is the total energy of a perfect supercell of Ni atoms and n is the total number of atoms in the system. Utilizing the above definition, a positive value of binding energy indicates a repulsive force between the two doping atoms, while a negative value indicates an attractive interaction. In this work, all of the binding energies are calculated by Eq. (3).

3. Result and discussion

In the previous work, Mottura et al.[11] concluded that magnetism is expected to have little effect on the binding energies of Re–Re, Ta–Ta, W–W pairs in the fcc-Ni lattice. To verify this conclusion, the binding energies between Re–Re, W–W, and Ta–Ta pairs are firstly calculated and the data are given in Table 2 and plotted in Fig. 2. The result based on non-spin-polarized calculations is consistent with that given by Mottura et al.[11] However, our results show that spin-polarized calculations and non-spin-polarized calculations will result in different conclusions for the Re–Re pairs. Therefore, the effect of magnetism on the binding energies of a XY nearest-neighbor pair is studied by comparing the spin-polarized and non-spin-polarized calculations.

Fig. 2. (color online) The binding energies between Re–Re, W–W, and Ta–Ta pairs in a 3 × 3 × 3 fcc Ni supercell containing 108 atoms as a function of XX separate distance: (a) spin-polarized calculations, (b) non-spin-polarized calculations.
Table 2.

The binding energies (in eV) between Re–Re, W–W, and Ta–Ta pairs at different distances in single-crystal nickel-based superalloys.

.

As shown in Table 1, the results show positive binding energies for Re–Re 1st NN and 2nd NN pairs by non-spin-polarized calculations, but negative binding energies by spin-polarized calculations, and the 2nd NN pair is energetically more stable. The 2nd NN rhenium pair is energetically stable at 0 K, suggesting that rhenium clusters would not form in the γ phase of single-crystal nickel-based superalloys. Therefore, it is expected that Re will approach a random distribution when alloyed in the γ phase with the concentration of 1.0–2.0 at.% typically used in these materials. Meanwhile, the results of W–W pairs and Ta–Ta pairs are almost the same in both spin-polarized and no-spin-polarized calculations. The repulsion energies observed for the Ta–Ta and W–W nearest neighbor pairs are rapidly reduced as the atoms are placed further apart, suggesting that the interactions between these solute atoms are strongly localized. It is possible to form short-range ordered structures and a Re–Re periodic chain along the ⟨100⟩ direction. Luo et al.[28] reported that heavy-atom columns align in the ⟨100⟩ and ⟨110⟩ directions and form heavy-atom stripes of about 1–2 nm in length. Heavy-atom ordering resembles the results reported by Maisel et al.,[29] which suggested that Re prefers to align in the ⟨420⟩ direction in the form of Ni4Re. The Ni4Re compound was shown to be stable by quantum mechanical high-throughput calculations at 0 K. Owing to the high diffusion activation energy of the heavy atoms and the stronger bonding strength of the refractory element with Ni, this short-range ordering will help to improve the mechanical performance.

The binding energies between X(Re, Ru)–Y(Re, W, Ta, Ru, Mo, Co, Cr, Al) pairs in Ni106XY system vs. XY separate distance are shown in Fig. 3. The binding energies essentially decay to zero when the distances between XY pairs increase. The interactions are only significant through the 3rd NN, which is not surprising as in a metallic alloy, the charges on the ion cores are screened strongly by the conduction free electrons.

Fig. 3. The binding energies between X(Re, Ru)–Y(Re, W, Ta, Ru, Mo, Co, Cr, Al) pairs in Ni106XY system vs. XY separate distance: (a), (b) spin-polarized calculation; (c), (d) non-spin-polarized calculation. Ranges from the 1st NN to 3rd NN distances of pure Ni are indicated by the dashed vertical lines.

We next discuss the effect of magnetism on the binding energies of an XY 1st NN pair by comparing the spin-polarized and non-spin-polarized calculations. The results are plotted in Fig. 4. It is shown that the binding energy strongly depends on the spin polarization. For example, under the spin-polarized calculations, the pairs of Re–Re, Re–Cr, Ru–Cr, Re–Ru, Ru–Mo, Ru–W, Re–Mo, and Mo–Cr at 1st NN have big attractive binding energies; while the pairs of Co–Cr, Co–Ru, Co–Mo, Co–W, Co–Re at 1st NN have a weak chemical repulsive tendency. The results of the non-spin polarization calculation are opposite to those of the spin polarization calculation. Under non-spin polarization conditions, most of pairs have big repulsive binding energies and are strongly repulsive in the first neighbor. Short-range order does not form, and it is randomly distributed over long-range. It is worth noted the results of Re–Cr, Ru–Cr, W–Cr, Ta–Cr, Re–Al, W–Al, Ta–Al, and Ru–Al with spin polarization are consistent with the experimental and calculation results reported by Huang et al.[19] The result of Re–Ru with spin polarization is consistent with the previous calculation result and experimental results reported by our research group.[18] Based on the existing experimental results, we believe that the results of spin polarized calcuation are more convincing. Therefore the following discussion are mainly based on the results of spin polarized calcuation, meanwhile the data of nonspin polarized calcuation are given to understand the magnetic effect.

Fig. 4. The binding energies between X(Re, W, Ta, Ru, Mo, Co, Cr, Al)–Y(Re, W, Ta, Ru, Mo, Co, Cr, Al) pairs in Ni106XY system at 1st NN distance. The blue histogram represents the result of spin-polarized calculation, and the red histogram represents the result of non-spin-polarized calculation.

The detailed binding energies between X (Re, W, Ta, Ru, Mo, Co, Cr, Al) and Y (Re, W, Ta, Ru, Mo, Co, Cr, Al) pairs with 1st NN, 2nd NN, and 11th NN distances are listed in Table 3. In the spin-polarized calculations, for the binding energies of pairs of Al–X, Ta–X, and W–X, almost all the solute elements exhibit alternating positive and negative binding energies from the 1st to 2nd NN sites, which indicates that all refractory elements are substituting on the Al-sublattice and form an L12 ordered structure with Ni, except for Co. The pairs of Re–Ru, Re–Cr, Ru–Cr, Re–Re, Mo–Cr, Re–Mo, and Ru–Mo have attractive binding energies at 1st and 2nd NN distances. The results show that Re–Re, Re–Cr, Mo–Cr, and Re–Mo at 2nd NN are more stable, Re–Ru and Ru–Cr at 1st NN are more stable, consistent with the previous experimental and calculation results.[18,19] The binding energies of Co-X pairs are close to 0, indicating that cobalt and other elements interact relatively weakly. We believe that cobalt and other elements do not have preference of occupancy and are randomly distributed. From the above analysis of the binding energies, the metals can be divided into three classes: (i) Al, Ta, and W exhibit repulsive binding energies with second elements at the 1st NN distance, but they have attractive binding energies with the second element at the 2nd NN distance; (ii) Re, Mo, Cr, and Ru have an affinity with each other and attractive binding energies at 1st NN and 2nd NN distances; and (iii) Co interacts with other elements relatively weakly.

Table 3.

The binding energies (in eV) between X(Re, W, Ta, Ru, Mo, Co, Cr, Al)–Y(Re, W, Ta, Ru, Mo, Co, Cr, Al) pairs in Ni106XY system at 1st NN, 2nd NN, and 11th NN distances.

.

Figure 5 shows the binding energies between Re and the other 28 elements. It is not difficult to find that the binding energy has an obvious periodicity, and different groups show different rules. In the spin-polarized calculations (see Fig. 5(a)), the 1st NN binding energy of Re–X exhibits a V-shaped curve, the lowest binding energy is located at half-full d = 5 and the value at the valley is less than 0. We believe that Re and other elements almost all have the 1st NN repulsive and 2nd NN attracting interaction energies, except that the pairs of Re–Re, Re–Cr, Re–Tc, Re–Ru, and Re–Os are attractive both at the 1st NN and 2nd NN. The binding energise of Re–Ru and Re–Os at 1st NN are lower than those at 2nd NN. It is possible to have short-range ordered structures in the ⟨110⟩ and ⟨100⟩ directions. Forming a Re-X periodic chain along the ⟨110⟩ direction may reflect the strengthening effect of Re. On the other hand, in the non-spin polarized calculation shown in Fig. 5(b), the 1st NN binding energy also exhibits a V-shaped curve, the lowest binding energy is located at d = 7 and the value at the valley is greater than 0, except for element Co. The binding energies at the second nearest neighbor are almost greater than 0, except that some early transiton metal elements (Sc, Ti, Y, Zr, Hf ) at the 2nd NN have a weak attraction. The binding energies of Re–Co at the 1st NN and 2nd NN distances all approach 0, indicating that there is a random distribution between Re and Co.

Fig. 5. The binding energies calculated with spinpolarization (a) and without spinpolarization (b) between X(Re)–Y(3d, 4d, 5d transition metal and Al) pairs in Ni106XY system at 1st and 2nd distances.

In order to systematically study the interactions between elements in the periodic table and to explore the rules from the data, we calculate the binding energies between combinations of the 28 elements. The binding energies of XY pairs in Ni106XY system at 1st NN by spin polarized calculation and non-spin polarized calculation are shown in Figs. 6 and 7, respectively.

Fig. 6. The binding energies between X(3d, 4d, 5d transition metals and Al)–Y(3d, 4d, 5d transition metals and Al) pairs in Ni106XY system at 1st NN distance by spin polarized calculation.
Fig. 7. The binding energies between X(3d, 4d, 5d transition metals and Al)–Y(3d, 4d, 5d transition metals and Al) pairs in Ni106XY system at 1st NN distance by non-spin polarized calculation.

In Figs. 6 and 7, the circle represents the positive binding energy, the triangle represents the negative binding energy, the box represents the binding energy close to zero. The deeper red color represents the greater binding energy, the darker blue color represents the more negative binding energy, the lighter color indicates that the binding energy is closer to zero. From Fig. 6, it can be found that the binding energies between the two elements have an obvious periodicity. The binding energy is closely related the electronic configuration of the elements. The behaviors of different elemental periods are not the same, the properties of 4d and 5d group elements are similar. For the early transition elements (Sc, Ti, V, Y, Zr, Nb, Hf, Ta), the sum of the d-electrons of the two element combinations is less than 6 and it exhibits repulsive effects. From the map, it is found that the triangular area forms a band, especially when the d electronic shell is fully filled (Xd + Yd = 9, 10, 11), such as Re+Cr, Re+Re, Re+Ru, Ru+Cr, etc. Its magnitude will decrease rapidly with decreasing overlap of the d shells of the adjacent atoms. Its magnitude will therefore be smallest for those elements with Xd + Yd nearly fully filled: (i) Xd + Yd = 2, 3, 4, 5, 6, 7, 8, repulsive binding energies for combinations of early transition metals; (ii) Xd + Yd = 9, 10, 11, attractive binding energies when the d-electrons of the element close to half full-shell state possess attractive binding energies. These results reflect the effect of the Hund’s rule.

This law may be related to the spin coupling between the incomplete d shells and the conduction electrons which leads to a tendency for a ferromagnetic alignment of the d electrons. Specific deep-level physics mechanism will be in-depth studied in future research. Additionally, further work studying the effects of alloy elements from electronic structure will always be of great importance.

4. Summary

The site preferences of two alloying elements XY in γ-Ni of Ni-based superalloy are systematically studied using first-principles calculations with and without spin-polarization. The doping elements X and Y are chosen from the 27 kinds of 3d, 4d, 5d group transition metals (Sc, Ti, V, Cr, Mn, Fe, Co, Cu, Zn, Y, Zr, Nb, Mo, Tc, Ru, Rh, Pd, Ag, Cd, Hf, Ta, W, Re, Os, Ir, Pt, Au) and Al. We find that the spin-polarized calculations for Re–Re, Re–Ru, Re–Cr, Ru–Cr show a strong chemical binding affinity between the solute elements and more consistent with the experimental results. The binding energies of pairs between the 28 elements have an obvious periodicity and are closely related the electronic configuration of the elements. When the d-electrons of the element is close to the half full-shell state, two alloying elements possess attractive binding energies, this result reflects the effect of the Hund’s rule. The combination of early transition metals (Sc, Ti, V, Y, Zr, Nb, Hf, Ta) have a repulsive interaction in γ-Ni. These results offer insights into the role of alloying elements in strengthening mechanism of superalloy.

Reference
[1] Reed R C 2006 The Superalloys: Fundamentals and Applications Cambridge Cambridge University Press 10.1017/CBO9780511541285
[2] Pollock T M Tin S 2006 J. Propulsion Power 22 361
[3] Blavette D Caron P Khan T 1986 Scr. Metall. 20 1395
[4] Blavette D Caron P Khan T 1988 Sixth Int. Symp. Superalloys-Superalloys 1988 305
[5] Blavette D Cadel E Pareige C Deconihout B Caron P 2007 Microscopy Microanalysis 13 464
[6] Wanderka N Glatzel U 1995 Mater. Sci. & Eng. A (Structural Mater.: Properties Microstruct. Process.) A203 69
[7] Rüsing J Wanderka N Czubayko U Naundorf V Mukherji D Rösler J 2002 Scr. Mater. 46 235
[8] Mottura A Miller M K Reed R C 2008 The 11th International Symposium on Superalloys-Superalloys 2008 Champion, Pennsylvania 2008 891
[9] Mottura A Warnken M Miller M K Finnis M W Reed R C 2010 Acta Mater. 58 931
[10] Mottura A Wu R T Finnis M W Reed R C 2008 Acta Mater. 56 2669
[11] Mottura A Finnis M W Reed R C 2012 Acta Mater. 60 2866
[12] Zhu T Wang C Y Gan Y 2010 Acta Mater. 58 2045
[13] Huang M Zhu J 2016 Rare Met. 35 127
[14] Smith J 1987 Effects of Cr and Re additions upon coarsening and deformation behavior of single-crystal Ni-base model superalloys Ph. D. Dissertation Illinois Illinois University Urbana
[15] Chen J Y Feng Q Cao L M Sun Z Q 2011 Mater. Sci. Eng. 528 3791
[16] Ge B H Luo Y S Li J R Zhu J 2010 Scr. Mater. 63 969
[17] Huang M Cheng Z Y Xiong J C Li J R Hu J Q Liu Z L Zhu J 2014 Acta Mater. 76 294
[18] Yu X X Wang C Y Zhang X N Yan P Zhang Z 2014 J. Alloys Compd. 582 299
[19] Huang Y Y Mao Z G Noebe R D Seidman D N 2016 Acta Mater. 121 288
[20] Kresse G Hafner J 1993 Phys. Rev. 48 13115
[21] Kresse G Hafner J 1994 Phys. Rev. 49 14251
[22] Kresse G Furthmüller J 1996 Phys. Rev. 54 11169
[23] Kresse G Furthmüller J 1996 Comput. Mater. Sci. 6 15
[24] Kresse G Joubert D 1999 Phys. Rev. 59 1758
[25] Perdew J P Burke K Ernzerhof M 1996 Phys. Rev. Lett. 77 3865
[26] Monkhorst H J Pack J D 1976 Phys. Rev. 13 5188
[27] Methfessel M Paxton A T 1989 Phys. Rev. 40 3616
[28] Luo Y S Zhang L H Wang Y M Ge B H Guo W Zhan J Zhang J X Zhu J 2016 Philos. Mag. Lett. 96 432
[29] Maisel S B Schindzielorz N Mottura A Reed R C Müller S 2014 Phys. Rev. 90 094110