Minru Wena and
Chong-Yu Wang*ab
aDepartment of Physics, Tsinghua University, Beijing 100084, China. E-mail: cywang@mail.tsinghua.edu.cn
bCentral Iron and Steel Research Institute, Beijing 100081, China
First published on 11th August 2016
The ideal strength, which is the minimum stress required to yield a perfect crystal, is fundamental for understanding the mechanical properties of real materials. Using density functional theory, the effect of the alloying elements Ta, W, Re, Mo, and Co on the ideal tensile strength (σIT) of γ′-Ni3Al along three characteristic directions was investigated. By examining the crystal mechanical stability in terms of Born criteria at every single strain during the tensile tests, the σIT values of L12-Ni3Al along the [001], [110] and [111] directions are determined to be 12.93 GPa, 7.20 GPa and 25.55 GPa, respectively. The effect of alloying elements on σIT shows obvious directionality because of directional interactions between impurities and host atoms. Doping with Ta, W, Re, Mo, and Co increases σIT in the [110] direction by 13%, 18%, 21%, 17%, and 5%, respectively. Furthermore, the electronic mechanism underlying the strengthening effect of alloying elements is determined by analyzing the d-orbital partial density of states and charge redistribution.
With advances in computational methods and modern computers, the ideal strength can be obtained by first-principles calculations.2,5–7 Early studies focused on the ideal strength of perfect crystals, such as single crystals (e.g., Al, Cu, Fe, and Ni)2,8–12 and ordered compound systems (e.g., FeAl, Co3(Al,W), and SiN).13–15 Calculation of the ideal strength has attracted great interest because it plays a crucial role in research of superhard synthesized materials.16–19 Furthermore, the ideal strength test has been extended to defective systems. For example, Hu et al.20 and Xu et al.21 studied the ideal strength of intermetallic compounds with a point defect, Lu et al.22 investigated the ideal tensile strength (σIT) of the Al grain boundary, and Yin et al.23 investigated the tensile and fracture process of the TiN/VN interface. Additionally, the temperature-dependent ideal strength was investigated by Shang et al.24,25 under the assumption that the effect of temperature only causes volume variation and electron–phonon and phonon–phonon interactions can be ignored.
Nickel-based single-crystal superalloys exhibit excellent elevated-temperature mechanical properties and are widely used for blade applications in advanced aeroplane engines and power-generation turbines.26,27 In general, the γ′-Ni3Al precipitate phase with its anomalous temperature dependence of the yield stress is largely responsible for the mechanical strength of alloys and their resistance to deformation at high temperature.28–30 Therefore, a study of the mechanical properties of γ′-Ni3Al should partially reflect the global properties of nickel-based superalloys. By adding alloying elements (e.g., Ta, W, Ti, Co, Cr, Re, Mo, and Ru), modern commercial nickel-based superalloys achieve extraordinary high-temperature performance.26,27,31,32 Ab initio calculations have been extensively used to investigate the influence of alloying elements on the properties of γ′-Ni3Al.33–37 However, these studies generally focus on the site preference of refractory elements in γ′-Ni3Al. The ideal strength is not emphasized. In particular, no theoretical study has focused on the effect of W, Ta, Mo, and Co on the ideal tensile strength of nickel-based superalloys, although these alloying elements have a significant influence on the properties of alloys.38–40
In this study, we performed first-principles calculations to investigate the effect of alloying elements (Ta, W, Re, Mo, and Co) on σIT of γ′-Ni3Al along the [001], [110] and [111] directions. The bonding nature of the alloying elements with host atoms, lattice evolution, and charge redistribution in the tensile process were analyzed.
![]() | (1) |
Based on the hypothesis of a dilute alloy, we used a single-impurity model41 to explore the strengthening effects of the alloying elements. Previous studies have shown that Ta, W, Re, and Mo prefer the Al site, whereas Co changes its site preference depending on the alloying composition in the γ′ phase.34,36,42,43 Therefore, the model with alloying element X (X = Ta, W, Re, Mo, and Co) substituting the Al atom was used. A 64-atom 2[10]a0 × 2[110]a0 × 2[001]a0 supercell was used for the σIT test of the [001] and [110] directions, and a 48-atom 2[1
0]a0 × [111]a0 × [11
]a0 supercell was used to calculate σIT in the [111] direction, as shown in Fig. 1(a) and (b). The sizes of these two supercells can avoid interactions between doping atoms.
We used a total-energy method based on density functional theory44,45 implemented in the Vienna Ab initio Simulation Package code.46 The ion–electron interaction was described by the projector augmented wave method,47 and the generalized gradient approximation of Perdew et al.48 was used to describe the exchange–correction function. The plane wave cutoff was set to 350 eV. Following the Monkhorst–Pack scheme,49 5 × 5 × 7 and 8 × 5 × 6 k-point meshes were used for tensile tests of the 64-atom and 48-atom models, respectively. The energy convergence of electronic self-consistency was on the order of 10−5 eV.
Table 1 lists the calculated lattice parameter a0 and elastic properties of L12-Ni3Al along with the corresponding values from experiments and other DFT calculations. Furthermore, the effects of alloying elements on the elastic properties of γ′-Ni3Al are also tabulated in Table 2. A 32-atom 2[001]a0 × 2[010]a0 × 2[100]a0 supercell and a 9 × 9 × 9 k-points mesh are used for the calculations of the elastic constants of alloying systems. The lattice parameter a0 and bulk modulus B are obtained by fitting the energy and volume data to the Murnaghan equation of state.50 Under small strain, the elastic energy density of a solid is given by:51
![]() | (2) |
Cal.a | Exptb | Exptc | Cal.d | Cal.e | Cal.f | |
---|---|---|---|---|---|---|
a This work.b Ref. 55.c Ref. 56.d Ref. 14.e Ref. 36.f Ref. 37. | ||||||
a0 (Å) | 3.567 | 3.57 | 3.57 | 3.568 | 3.579 | 3.56 |
B (GPa) | 175.9 | 173 | 174.3 | 174.9 | 180.4 | 185 ± 6 |
C11 (GPa) | 227.9 | 223 | 227 | 229.7 | 243.8 | 246 ± 6 |
C12 (GPa) | 149.9 | 148 | 148 | 147.5 | 148.7 | 155 ± 6 |
C44 (GPa) | 120.9 | 125 | 120 | 116.6 | 123.4 | 130 ± 0.0 |
G (GPa) | 76.9 | 77 | 77 | 76.8 | 84.2 | 96 ± 0.0 |
E (GPa) | 201.5 | 202 | 201 | 201 | 218.5 | |
G/B | 0.44 | 0.45 | 0.44 | 0.44 | 0.47 | |
ν | 0.309 | 0.308 | 0.308 |
Ni3(Al,Ta) | Ni3(Al,W) | Ni3(Al,Re) | Ni3(Al,Mo) | Ni3(Al,Co) | Ni3Al | |
---|---|---|---|---|---|---|
a0 (Å) | 3.579 | 3.573 | 3.571 | 3.573 | 3.559 | 3.567 |
B (GPa) | 180.1 | 182.5 | 183.2 | 181.3 | 179.0 | 175.9 |
C11 (GPa) | 237.5 | 241.9 | 243.9 | 239.7 | 231.0 | 227.9 |
C12 (GPa) | 151.4 | 152.7 | 152.8 | 152.0 | 152.9 | 149.9 |
C44 (GPa) | 122.5 | 126.7 | 127.9 | 124.6 | 118.1 | 120.9 |
G (GPa) | 80.6 | 83.4 | 84.6 | 82.0 | 75.8 | 76.9 |
E (GPa) | 210.4 | 217.1 | 220.0 | 213.8 | 199.4 | 201.4 |
G/B | 0.45 | 0.46 | 0.46 | 0.45 | 0.42 | 0.44 |
ν | 0.305 | 0.302 | 0.300 | 0.303 | 0.314 | 0.309 |
Generally, the ideal strength of crystals refers to the first maximum point of the stress–strain curve. The corresponding stress at this point for a material is the maximum ideal tensile strength (σMIT). However, the lattice may become mechanically unstable prior to reaching the stress–strain peak in terms of violating the Born criteria of stability under uniaxial load.51,57 It is important to examine the mechanical stability of a defect-free lattice during the tensile tests as previously verified.58 The cubic lattice becomes tetragonal, orthorhombic and trigonal under the [001], [110] and [111] directional loading, respectively. The mechanical stability criteria of tetragonal, orthorhombic and trigonal lattices are given by51,52,59
![]() | (3) |
C11 > 0, C22 > 0, C33 > 0, C44 > 0, C55 > 0, C66 > 0, (C11 + C22 − 2C12) > 0, (C11 + C33 − 2C13) > 0 and (C22 + C33 − 2C23) > 0, | (4) |
C11 > |C12|, (C11 + C12)C33 − 2C132 > 0 and (C11 − C12)C44 − 2C142 > 0, | (5) |
In terms of the [111] direction, the tensile stresses of doped Ni3Al also yield at a lower strain, as shown in Fig. 2(c). The lattice becomes mechanically unstable at a strain of 19% while the σMIT of Ni3Al is reached at a strain of 22% according to Fig. 3(c). Therefore, the ideal tensile strength of pure Ni3Al is 25.55 GPa in the [111] direction. Here, it is important to state that the supercell as shown in Fig. 1(b) was employed and treated as an orthorhombic lattice to calculate the elastic constants during the [111] tension. Certainly, the supercell still possesses six independent elastic constants, Cij, as shown in Fig. 3(c). Unlike the failure in the [001] tension, the mode of failure for the [111] tension is C44 = 0. Furthermore, the BCT stability zone (the mechanical stability zone in terms of Born criteria) includes the peak point from the σ–ε curve for alloying systems (as shown in Fig. 2(c)). Therefore, the σIT of the substitutional systems slightly decrease in the [111] direction. Taking into consideration the Born criteria of stability during uniaxial load, Table 4 lists the σIT values and corresponding strains in all of the calculated directions. Note that the calculated σMIT of non-doped γ′-Ni3Al is in agreement with the results of Wang et al.61 who observed that the addition of Re can improve σMIT of γ′-Ni3Al by 1% and 6% in the [001] and [111] directions, respectively. The different models used in their study and the present work may be responsible for the discrepancy. Here, we used single-impurity models and the distance between impurities is farther than the fifth nearest neighbor. Such supercells can avoid interactions between doping atoms during tensile tests. Furthermore, the stress–strain curves in Fig. 2(a)–(c) are continuous and smooth, which indicates the reliability of the method used in this study.
System | [001] | [110] | [111] | |||
---|---|---|---|---|---|---|
σIT (GPa) | ε | σIT (GPa) | ε | σIT (GPa) | ε | |
Ni3(Al,Ta) | 13.42 | 0.11 | 8.14 | 0.08 | 23.89 | 0.16 |
Ni3(Al,W) | 13.64 | 0.11 | 8.52 | 0.09 | 23.61 | 0.16 |
Ni3(Al,Re) | 13.71 | 0.11 | 8.73 | 0.09 | 23.2 | 0.16 |
Ni3(Al,Mo) | 13.28 | 0.11 | 8.43 | 0.09 | 23.64 | 0.16 |
Ni3(Al,Co) | 12.98 | 0.11 | 7.57 | 0.08 | 22.53 | 0.14 |
Ni3Al | 12.93 | 0.11 | 7.20 | 0.08 | 25.55 | 0.19 |
The stress–strain relationship in the [110] direction is different from those in the [001] and [111] directions. As shown in Fig. 2(a)–(c), the [001] and [111] directions can be considered as strong orientations with σIT as large as 12 GPa. However, the [110] direction is obviously the weakest direction for γ′-Ni3Al with σIT yielding only 7.2 GPa. In the C11 + C33 = 2C13 mode (as shown in Fig. 3(b)), the lattice remains true to the Born criteria of stability until a strain of 9%, and thereafter fails upon increasing strain in the [110] tension for non-doped, Ta-doped and Co-doped systems. For Ni3(Al,W), Ni3(Al,Re) and Ni3(Al,Mo), the point at which mechanical instability occurs is immediately prior to the yield point. Addition of alloying elements X (X = Ta, W, Re, and Mo) can considerably improve the tensile stress and σIT in the [110] direction, and Re exhibits the greatest improvement, as shown in Fig. 2(b). Note that doping with Co has a negligible effect on the stress of the γ′-Ni3Al phase. As listed in Table 4, the ability of the alloying elements to increase the σIT are in the order: Re > W > Mo > Ta > Co. Compared with pure Ni3Al, the σMIT of Ni3(Al,Re) improves from 7.2 to 8.75 GPa, an increase of 22%. Wang et al.61 found that addition of Re can increase σIT by 51% in the [110] direction. This larger increase is because of the higher alloying element concentration in their model.
According to the above results in relation to the mechanical stability in terms of Born criteria examinations, the σIT of L12-Ni3Al along the [001], [110] and [111] directions are determined to be 12.93 GPa, 7.20 GPa and 25.55 GPa, respectively. The [110] orientation is the weakest uniaxial tensile direction. It is important to mention here that we do not perform phonon calculations to verify the phonon stability during the tensile processes. In fact, phonon-induced instabilities may occur significantly prior to the stress–strain peak, similar to the scenario undergoing during the tension and 〈112〉{111} shear for an Al crystal.8 Consequently, our predictions may overestimate the ideal strength.
The effect of alloying elements on σIT of Ni3Al exhibits obvious directionality. The orientation dependence of the strengthening mechanism arises from interatomic interactions that vary significantly in different directions in Ni3Al. Analysis of charge redistribution will reveal these interactions between impurities and host atoms. Impurities have the strongest effect on the [110] direction. Hence, we will highlight the electronic mechanism of alloying elements during [110] tensile deformation in later analyses.
The dependence of the lattice parameters under [110] uniaxial strain is shown in Fig. 4. With uniformly increasing tensile strain (i.e., increasing lattice parameter a), c decreases and b remains almost constant. When the supercell of pure Ni3Al is deformed by 18%, the c lattice parameter is equal to half of a (see Fig. 4). Therefore, the distance between an Al atom and the first nearest neighbor (FNN) Ni atom along the [110] direction is equivalent to the distance between nearest Al atoms along the [001] direction, as shown in the inset of Fig. 5. Accordingly, at this point, the response of the energy under [110] tension has a saddle point, as shown in Fig. 5, which indicates that the new triclinic structure is an unstable “phase”. Consequently, the stress reaches the second zero-stress point according to eqn (1) and attains its maximum at an early strain of about half of 18%, which is consistent with the stress–strain relationship shown in Fig. 2(b). In contrast to the [110] loading direction, the energies monotonically increase between ε = 0.00 and ε = 0.50 under tension along the [001] and [111] directions (Fig. 5). It can be concluded that the σMIT of the [110] tension being the lowest σMIT among the three characteristic directional loading is an intrinsic characteristic of the L12 structure (space group: Pmm).
![]() | ||
Fig. 4 Variation of the lattice parameters of Ni3(Al,X) during the [110] tensile test. X = Ta, W, Re, Mo, Co, and Al, where X = Al represents non-doped Ni3Al. |
Fig. 4 also shows the influence of alloying elements on variation of the lattice parameters during the [110] tensile test. Doping with Co has a negligible effect on variation of the lattice parameters. Fig. 4 shows that Ta, W, Re, and Mo can considerably decrease the rate of variation of the b and c lattice parameters of γ′-Ni3Al. It is expected that Ta, W, Re, and Mo can improve the mechanical strength and deformation resistance of alloys. These results are consistent with the stress–strain relationships (Fig. 2(b)), and agree well with previous experiments.38,40,62,63
Although Co has a negligible effect on the stress of γ′-Ni3Al, the other doping elements (i.e., Ta, W, Re, and Mo) can enhance σIT in the weakest tensile direction. To reveal the strengthening mechanism, we investigated the density of states (DOS) of these alloying systems. Fig. 6(a)–(e) show the d-orbital partial DOS of doping element X and its FNN Ni atom along the [110] direction. The partial DOS of pure Ni3Al are also plotted in Fig. 6(f) for comparison. As shown in Fig. 6(a)–(e), all of the alloying elements except Co have a prominent feature: the noticeable bonding and antibonding region of X-d (X = Ta, W, Re, and Mo) orbitals is well separated by a deep valley near the Fermi level. This feature indicates that these alloying elements can stabilize the system by substituting for the Al atom in γ′-Ni3Al.64 Additionally, there are strong interactions between X and the host atom Ni in the doping systems by d–d hybridization (see Fig. 6). Here, we indicate the resonant hybridization peaks by dotted lines in Fig. 6(a)–(f). Interaction between Re and Ni shows the greatest hybridization, followed by W–Ni, Mo–Ni, and Ta–Ni. The partial DOS of Ni3(Al,Co) behaves in a different way: the Fermi level is located at the peak of the Co-d DOS and the interaction between Co-d and Ni-d is weak. It is expected that Co will improve the ductility of Ni3Al rather than the mechanical strength. Consequently, addition of Co has a negligible effect on the response of the tensile stress to strain during the [110] deformation test (see Fig. 2(b)).
We next investigated the charge density difference (Δρ) of Ni3(Al,X) in the process of tensile deformation. Charge redistribution induced by impurities can directly reflect the bonding nature between alloying atoms and host atoms. The Δρ between the impurity atom and its FNN host atoms (twelve Ni atoms) in real space and Δρ along the [110] direction in [110] tensile deformation are shown in Fig. 7. Electrons flow from both Ni and X (X = Ta, W, Re, and Mo) atoms into covalent-like Ni–X bonds (Fig. 7(a′)–(d′)). Charge accumulation shows robust directionality, mainly appearing along the 〈110〉 direction for X-doped systems (X = Ta, W, Re, and Mo), as shown in Fig. 7(a)–(d). This suggests that interactions between these impurities and Ni atoms exhibit strong directionality, which is the origin of their directional strengthening effect on the σIT in the γ′-Ni3Al phase.
Moreover, Δρ is quite different dependent upon the alloying system. As shown in Fig. 7(b) and (c), Δρ in Re-doped and W-doped systems exhibit similar features. However, although Ta, W and Re are the adjacent element in the sixth period, Ta has a considerably weaker effect than W and Re (see Table 4) on improving σIT in the γ′-Ni3Al phase. Fig. 7(a–c) reveals the origin of the strengthening effect. Charge accumulation in the W-doped and Re-doped systems is stronger than in the Ta-doped system. Interaction of W–Ni and Re–Ni can effectively improve the mechanical strength, which agrees with experimental investigations.38,62 As shown in Fig. 7(d), Δρ in the Mo-doped system shows stronger charge accumulation than the Ta-doped system but weaker than the W-doped and Re-doped system, which is consistent with previous calculations of the ideal tensile stress (Table 4). Unlike the other alloying systems, the missing charge accumulation at the isovalue of 0.0672 e Å−3 in the Co-doped system (Fig. 7(e)) suggests more metallic bonding character between Co and FNN Ni atoms. It is expected that doping with Co can improve the malleability and ductility in nickel-based alloys rather than the mechanical strength. Al has an electronic structure of 3s23p1 with only three valence electrons, which greatly contributes to it having the smallest σIT among all of the investigated systems. Accordingly, compared with alloying elements, Δρ near the Al atom is very small in pure Ni3Al (see Fig. 7(f′)). In all of the systems, with increasing tensile strain, the distances between atoms along the [110] direction increases and Δρ becomes smoother, as shown in Fig. 7(a′)–(f′).
This journal is © The Royal Society of Chemistry 2016 |