† Corresponding author. E-mail:
Project supported by the National Basic Research Program of China (Grant No. 2011CB606402).
The effects of boron and carbon on the structural, elastic, and electronic properties of both Ni solution and Ni3Al intermetallics are investigated using first-principles calculations. The results agree well with theoretical and experimental data from previous studies and are analyzed based on the density of states and charge density. It is found that both boron and carbon are inclined to occupy the Ni-rich interstices in Ni3Al, which gives rise to a cubic interstitial phase. In addition, the interstitial boron and carbon have different effects on the elastic moduli of Ni and Ni3Al. The calculation results for the G/B and Poisson’s ratios further demonstrate that interstitial boron and carbon can both reduce the brittleness of Ni, thereby increasing its ductility. Meanwhile, boron can also enhance the ductility of the Ni3Al while carbon hardly has an effect on its brittleness or ductility.
Ni-based superalloys have been successfully utilized in both land-based and aeroplane gas turbine industries because of their excellent oxidation resistance and favorable high-temperature properties, including strength, fracture toughness, ductility, fatigue resistance, and enhanced creep. These desirable properties generally result from the incorporation of coherent, ordered γ′-Ni3Al phases into the disordered, γ-Ni solid solution matrix.[1–4] For superalloys, elastic properties are undoubtedly essential. Thus, much theoretical and experimental research has concentrated on the elastic properties of the γ-Ni and γ′-Ni3Al phases of Ni-based superalloys.[5–18]
Wang Y J and Wang C Y[10] investigated the effects of alloying elements (Ta, Mo, W, Cr, Re, Ru, Co, and Ir) on the elastic constants and elastic moduli of both Ni and Ni3Al using ab initio calculations. The results showed that all the elements considered increase the Young’s modulus of Ni. In contrast, only Re was found to be effective in increasing the shear modulus. Furthermore, all the alloying elements except Co increase the modulus of Ni3Al slightly. However, increasing the modulus leads to reduced Poisson’s ratio, so both Ni and Ni3Al might become more brittle with the increase of alloying element (except Co). This restricts the applicabilities of Ni-based superalloys at high temperatures.
It is universally received that the ductility of polycrystalline Ni3Al at ambient temperature can be greatly improved by adding a small amount of boron.[19] This ‘boron effect’ has been confirmed in cast and recrystallized,[20] and rapidly solidified[21] Ni3Al. Several mechanisms have been proposed to explain the boron effect on the ductility of Ni3Al alloy.[20,22–25] Liu et al. suggested that boron increases the cohesive strength of the grain boundary.[20] George et al. argued that boron suppresses moisture-induced embrittlement either by reducing the rate of hydrogen diffusion in Ni3Al or by lowering the segregation of hydrogen atoms at the grain boundaries.[25] Besides, Huang et al.[26] found that boron also exhibits a large solid-solution strengthening effect in rapidly solidified Ni3Al. This large strengthening potency of boron is possibly associated with the large lattice strain induced by occupying interstitial lattice positions,[26] which is in accordance with the theory of Mott and Nabarro.[27] The single-crystal tensile tests by Heredia and Pope,[28,29] however, clearly show that boron also improves the ductility of the bulk material. Sun et al.[30] found that boron also increases the maximum cleavage stress (ideal yield stress) of single Ni3Al crystal. It was also suggested that a ‘bulk effect’ should be considered in addition to the grain-boundary strengthening effect due to boron while trying to explain the improvement in ductility of polycrystalline Ni3Al with boron addition.[30]
A first-principles method was employed to investigate the segregation behaviors of hydrogen and boron in Ni-based and Ni3Al-based alloys by Wu et al.[4] The analysis of the chemical binding energy showed that boron is able to segregate at the interstices in the Ni phase, Ni3Al phase, and Ni/Ni3Al interface. It was found that the addition of boron to both Ni-based and Ni3Al-based alloys can improve their ductilities and that this boron-induced ductility can be controlled and improved by manipulating the lattice misfit.
Another interstitial element, carbon, may also affect the properties of Ni3Al. It is surprising that carbon in Ni3Al, which strongly segregates grain boundaries, can, in fact, promote intergranular brittleness (contrary to boron behavior), but still exert a large strengthening effect on the solid solution, which is similar to boron’s.[31] This is also (probably) associated with the resulting increase in the lattice constant of Ni3Al caused by carbon intercalation.[32,33] Besides, carbon induces oxide cavities to migrate outwards via a dissociative mechanism assisted by the gaseous transfer of oxygen across the cavities wherein CO–CO2 acts as a carrier during high-temperature oxidation of Ni.[34] Most of the previous studies have mainly focused on the effects of boron or carbon on the properties of polycrystalline Ni or Ni3Al, so the available data about the effects of these two elements on the properties of single-crystal (SC) Ni and Ni3Al are very limited. Few reports have systematically studied the effects of boron and carbon on the elastic properties of SC Ni and Ni3Al.
Density functional theory (DFT) has played an important role in previous studies on the mechanical properties of materials.[35–37] In light of the interesting effects of boron and carbon, we therefore present here an ab initio investigation of the effects of these two elements on the structural, elastic, and electronic properties of both Ni solution and Ni3Al intermetallics. Such an addition is expected to improve the brittleness values of SC Ni and Ni3Al by adding these two elements so as to find the substitutes or supplementary elements for heavy metals. Enthalpies of formation can be calculated and used to indicate the phase stability and to determine the site preferences of these two elements in Ni3Al. Based on the elastic constants thus calculated, elastic and orientation-dependent elastic moduli may be obtained and analyzed in terms of density of states and charge density. The results derived can be proved useful in designing more advanced alloys.
Unit cells of face-centered cubic (fcc) Ni and L12Ni3Al are shown in Figs.
In order to explore the changes in the structures, elastic constants, and various elastic moduli of Ni solution and Ni3Al intermetallics when they are doped with interstitial atoms X (X = B, C), supercells measuring 2 × 2 × 2 are selected for calculation purposes. The chemical compositions of the selected supercells are therefore Ni32X and Ni24Al8X (X = B, C). Supercells of this size are selected according to the concentrations of the alloying elements occurring in calculation models established in relevant first-principles studies,[30,39–41] which makes the interaction among the adjacent impurity atoms negligible and gives rise to more obvious effects from the impurity itself. To more clearly investigate the effects of the addition of interstitial boron or carbon on the structure and elastic properties of Ni and Ni3Al, a single-impurity model is adopted to study the case where the two interstitial impurities are contained in Ni and Ni3Al, respectively. At the same time, similarly-sized supercells of Ni and Ni3Al containing no impurity atoms are also studied for comparison with the doped supercells.
Figures
All the calculations carried out in this research are performed using the DFT-based Vienna ab initio simulation package (VASP).[42–45] The pseudo-potentials utilized to describe the interactions between nuclear and extranuclear electrons are obtained using the projector-augmented wave (PAW) method.[46] The energy cut-offs when doping with boron and carbon are taken to be 415 eV and 520 eV, respectively, to ensure that the total energy of the calculated physical model is fully converged. The exchange-correlation potential is obtained using the generalized gradient approximation (GGA) method developed by Perdew, Burke, and Ernzerhof (PBE).[47] Brillouin-zone integration is performed using a Monkhorst–Pack k-point mesh.[48] Tests on the effect of mesh size on the calculated mechanical and energy quantities suggest that a k-point mesh of size 9 × 9 × 9 has favorable convergence effect. However, in order to ensure that the calculation results are accurate and the total energy of the system is fully converged, an 11 × 11 × 11 k-point mesh is adopted in this research. The convergence criteria for electronic self-consistency and ionic relaxation are set to be 10−5 and 10−4 eV, respectively.
In order to accurately obtain the equilibrium structure and equilibrium volume of each system, the initial structure of each system (including the cell parameters and internal configuration) is fully relaxed to begin with. Then, the fully-relaxed supercells are allowed to undergo sufficient relaxation in the sense of fixing volume and changing shape. In doing so, the shapes of the supercell in each system can be determined in the equilibrium state (i.e., the angles and ratios between the basic vectors of the supercell). Afterwards, the volumes of the supercells are reduced or increased to different degrees compared with that of the sufficiently relaxed structure. After this, full relaxation (in the sense of fixing volume and shape) and static calculations are performed on a series of supercells with different volumes so as to accurately obtain the total energy. Finally, the equilibrium volume V0 and bulk modulus B of each system are obtained by fitting the energy and volume data to the Murnaghan equation of state.[49] The structural parameters of the supercells are then obtained in the equilibrium state. Using this as a basis, the internal configuration of the supercell in each system is required to experience sufficient relaxation once again so as to obtain the final equilibrium structure. Then, a fully static calculation is carried out to finally give the total energy E0 in the equilibrium state.
Enthalpy of formation is an important index[50] for judging the phase stability of a system. It can be calculated by various means.[51] In the present research, an effective method is used to find the formation enthalpies by calculating the relative value of the equilibrium energy of the compound and the composition-weighted average of energy of each pure component in the fcc structure.[52,53] For a binary system, ApBq, the enthalpy of formation ΔHfcc(ApBq) can be formulated as:
In this work, we begin with a specific shape and find the strain tensor ε produced in the crystal by changing the initial basic vector of its supercell. Then, the total energy of a system to which a group of strains with different amplitudes are applied is calculated. After this, data fitting is performed (quadratic or higher order) on groups of data points describing the variation of energy density with strain amplitude, obtained using first-principles calculations. In this way, the quadratic coefficients of the crystals (i.e., the elastic constants Cij) or a linear combination of elastic constants can be obtained,[56] which can also be obtained using other similar methods.[7–9] The strain tensor ε has been defined and labeled by Voigt.[57] As only the non-rotational strain needs to be taken into account, the strain can be expressed as a symmetric tensor containing six independent components:
The initial basic vector
As a general rule, a crystal may be described using 21 independent elastic constants. For cubic crystal systems, however, the high symmetry reduces the number of independent Cij terms to 3, which we take to be C11, C12, and C44. To obtain a value for C44, a strain in the form ε = (0,0,0,δ,δ,δ) can be applied to the cubic crystal, where δ corresponds to the strain amplitude.[10] By substituting the strain into Eq. (
Similarly, when a strain in the form ε = (δ,δ, (1 + δ)− 2 − 1,0,0,0) is applied to the cubic crystal, the relationship between the energy density and strain amplitude is found to be
For the calculations performed in this research, effective values of the strain amplitude δ are required by using a series of amplitude values varying from −0.03 to 0.03 (step size 0.005). This range ensures the accuracy and reliability of the fitted results when polynomial fitting is performed on the series of data points obtained for the variation of energy density with strain amplitude.
The elastic constants are determined using first principles calculations. These constants are then used to derive the bulk modulus B and shear modulus G of the crystal. The Young’s modulus E and Poisson’s ratio ν can then be obtained using the derived values for the bulk and shear moduli.[58] The two theoretical research studies by Voigt[57] and Reuss[59] provide two effective methods of calculating approximate values for the elastic moduli from the elastic constants. Besides, they are used for calculating the upper and lower limits of actual elastic modulus, respectively. The Young’s modulus E, Poisson ratio ν, approximate bulk modulus B, and shear modulus G can thus be obtained using the Voigt and Reuss theories from the following formulas:[58]
Elastic moduli obtained by calculating elastic constants through using the first-principles method are macroscopic average values. However, actual crystals are anisotropic. To calculate the Young’s modulus E and shear modulus G for a specific orientation of the cubic crystal system, the following expressions can be adopted:[57,60]
The lattice parameters and coordinates of the ions inside the Ni and Ni3Al supercells doped with interstitial atom X (X = B, C) and the pure Ni and Ni3Al are fully relaxed and optimized using the aforementioned method. Table
As shown in Table
Using 2 × 2 × 2 supercells of pure Ni and Ni3Al as the control group, the elastic properties of Ni and Ni3Al doped with interstitial impurities are subsequently calculated. Table
As can be seen in Table
Ni and Ni3Al experience the volume expansion to different extents when doped with boron and carbon, and the volume changes in Ni and Ni3Al due to boron doping are greater than those caused by carbon doping in the same interstice. In a cubic crystal system, the lattice parameter a presents a variation trend similar to that of the volume. This is attributed to the fact that the atomic radius of boron is greater than that of carbon. In the tetragonal crystal systems, the lattice parameter c of Ni24Al8B2 is also greater than that of Ni24Al8C2. However, when it comes to the lattice parameter a, the former is smaller than the latter and it is even smaller than that of pure Ni3Al. This indicates that the volume increase in Ni24Al8B2 is a result of the enlarged parameter c.
The average energy of a system can measure the system degree of stability. From Table
The enthalpies of formation of the interstitial phases (Ni32B, Ni32C, Ni24Al8B1, Ni24Al8B2, Ni24Al8C1, and Ni24Al8C2) can be calculated using Eq. (
Figure
Above all, it is found that both boron and carbon incline to occupy interstice 1 in Ni3Al, thus giving rise to the formation of the interstitial phase of Ni3Al with cubic structure. Therefore, in the following part, we only need to take into account the two cubic interstitial phases Ni24Al8B1 and Ni24Al8C1 of Ni3Al. For simplicity and convenience, Ni24Al8B1 and Ni24Al8C1 will henceforth be written as Ni24Al8B and Ni24Al8C, respectively (i.e., the superscript ‘1’ is simply omitted from the chemical compositions for clarity).
According to the above discussion, by applying specific strains to the crystal, the relationship between the energy density and strain amplitude can be utilized to calculate the elastic constants Cij of the cubic crystal. The results of such calculations for pure Ni, Ni3Al, and their interstitial phases are summarized in Table
As can be seen in Table
The behaviors of the three independent elastic constants of the interstitial phases Ni24Al8B and Ni24Al8C in comparison to those of pure Ni3Al are more complex. As shown in Table
Above all, the presence of interstitial boron and carbon reduces all three elastic constants of Ni. In contrast, in the Ni3Al case, two of the elastic constants of the two interstitial phases of Ni3Al, C12 and C44, are clearly lower than those of pure Ni3Al, however C11 is slightly higher than that of the pure Ni3Al. Therefore, it can be speculated that the presence of interstitial boron and carbon will be expected to effectively reduce the various elastic moduli of Ni and Ni3Al, and thus increase the values of ductility and toughness of the Ni and Ni3Al.
As is well known, the mechanical stability is an important area of theoretical research when considering the phase stability of material. The mechanical stability of an alloy phase can be evaluated using the elastic constants of the corresponding SC.[69] The criteria for determining the mechanical stability of cubic crystal structure can be expressed as the following formulas:
It can be shown using the values in Table
The first-principles method is used to calculate a series of data points for the energy–volume of each system and the bulk modulus B of each system is found by fitting the data to the Murnaghan equation of state. The general relationships between the elastic constants and shear modulus (approximated by Voigt and Reuss) are shown in Eqs. (
The average value G = (GV + GR)/2
of the shear moduli calculated using the Voigt and Reuss approximations is adopted as the final shear modulus. The Young’s modulus E and Poisson’s ratio ν are calculated from Eqs. (
It can be seen from Table
On the one hand, interstitial boron and carbon both enhance the bulk modulus of Ni3Al, and the strengthening effect by carbon is more obvious than by boron. On the other hand, boron is expected to reduce the shear modulus, whereas carbon causes it to slightly increase. The results for shear modulus as calculated using the Voigt and Reuss approximations are also different: both boron and carbon reduce the Voigt-approximated shear modulus and boron presents a larger reduction effect. However, carbon significantly increases the value of the Reuss-approximated shear modulus, while boron hardly has any effect on the shear modulus in the Reuss approximation (which basically remains the same as that of pure Ni3Al). Finally, boron effectively reduces the Young’s modulus of Ni3Al while carbon enhances it. This variation is similar to that found in the shear modulus.
The ratio of the shear modulus to the bulk modulus (G/B) and Poisson’s ratio ν can favorably reflect the ductility or brittleness of a material: the smaller the value of G/B, or the greater the Poisson’s ratio ν, the more ductile the material is, and vice versa.[71] As demonstrated in Table
The Young’s moduli and shear moduli of Ni and Ni3Al before and after adding the boron and carbon are calculated using the averaging method proposed by Hill.[58] This method can favorably reflect the average and global properties of these materials. However, as an actual crystal is not isotropic, the Young’s modulus and shear modulus are dependent on the crystal’s orientation. Thus, they need to be formally investigated as a function of orientation.
The Young’s and shear moduli of the six cubic interstitial phases are subsequently investigated for certain crystal orientations using Eq. (
As can be seen from Fig.
For Ni3Al, adding the interstitial boron and carbon reduces the shear modulus in the [100] orientation and the Young’s modulus in the [111] orientation (the greater reduction effect by boron will occur). Both interstitial impurities increase the Young’s modulus of Ni3Al in the [100] orientation (the superior strengthening effect by carbon will be present). This behavior is similar to the variation of the Young’s modulus in the [100] orientation before and after adding the interstitial boron and carbon to Ni. However, the shear modulus variations in the [110] and [111] orientations and the Young’s modulus variation in the [110] orientation of Ni3Al before and after being doped with interstitial boron or carbon are different from those occurring in Ni. To be specific, both interstitial boron and carbon can enhance the shear modulus in the [110] and [111] orientations, and the strengthening effect of carbon is superior to that of boron. However, the two interstitial atoms exhibit contrary influences on the Young’s modulus in the [110] orientation: boron reduces it, while carbon enhances it. Meanwhile, it is found that the shear modulus in each of the systems always presents the following ascending order: [111] < [110]
< [100], while the Young’s modulus exhibits a descending order: [111] > [110]
> [100]. Additionally, it can also be found from Fig.
The variations of the elastic modulus in the different orientations before and after being doped with boron and carbon respectively indicate that the anisotropies of Ni and Ni3Al change as a result of doping. The anisotropy of crystal can be quantified using the anisotropic factor A and dimensionless parameter A*.[72] The former can be determined for cubic crystals using the three independent elastic constants as follows:
When A* has a value of 0, i.e., A = 1, the crystal is isotropic. When A < 1, A* increases as A decreases. When A > 1, A* increases with the increase of A. The value of A* is always positive, regardless of whether A < 1 or A > 1, and the larger the value of A*, the greater the degree of anisotropy in the crystal will be. The calculated values of these anisotropy parameters for Ni, Ni3Al, and the doped derivatives are illustrated in Fig.
It can be seen from Fig.
To further explore the reasons for the effects of doping on the properties of Ni and Ni3Al, the partial densities of states (PDOSs) of pure Ni, Ni3Al, and their interstitial phases containing boron and carbon are obtained by the first-principles method, respectively. The calculations are based on the optimized structures corresponding to each system. The results are shown in Fig.
In Fig.
It can also be found that after interstitial boron and carbon are added to Ni separately, the density of states of the 3d orbital extends towards the areas with low energy. This indicates that the 3d electrons in Ni migrate to the deep energy level. As a result, the energies of the Ni atoms that are closest to the impurity atoms are lower than those of the Ni atoms at the corresponding positions in pure Ni. This is the reason why the average energy of the interstitial phases of Ni is lower than that of pure Ni (as found in the aforementioned analysis). Additionally, it is found that the region of overlap between the density of states of the C-2p orbital and its NN Ni-3d orbital in Ni32C is larger than that of the B-2p orbital and Ni-3d orbital in Ni32B. This suggests that the interaction between a carbon atom and the Ni atoms is stronger than that between a boron atom and the Ni atoms. Therefore, the various elastic constants and moduli of Ni32C are larger than those of Ni32B.
For Ni24Al8B and Ni24Al8C, there is also strong hybridization between the interstitial boron and carbon atoms and their NN Ni atoms respectively, which gives rise to the stable doped structures. The strength of the hybridization between the C-2p and Ni-3d orbitals in Ni24Al8C is greater than that between B-2p and Ni-3d in Ni24Al8B. This implies that the interaction between C and Ni atoms is stronger than that between B and Ni, which also may be related to the various elastic constants and elastic moduli for different orientations of Ni24Al8C, which are higher than those of Ni24Al8B. It is also seen from Fig.
Differential charge density can be used to give a more visual representation of bonding conditions and charge transfer between atoms. Figure
It can be seen from Fig.
In this work, we systematically study the effects of interstitial boron and carbon on the structural, elastic, and electronic properties of Ni solution and Ni3Al intermetallics using the first-principles method. The results demonstrate that both boron and carbon can increase the lattice parameters and volumes of Ni and Ni3Al, and that they give rise to interstitial phases of Ni with cubic structures. However, as there are two different octahedral interstitial sites in Ni3Al, the interstitial phases of Ni3Al present cubic and tetragonal structures. By calculating the equilibrium energies and enthalpies of formation of each system, it is found that the presence of interstitial boron and carbon atoms can enhance the phase stabilities of Ni and Ni3Al. Furthermore, carbon has a more obvious strengthening effect than boron. In addition, both boron and carbon incline to occupy the Ni-rich interstice (interstice 1) in Ni3Al, which therefore results in the preferential generation of the cubic interstitial Ni3Al phase.
The elastic constants and various elastic moduli of pure Ni and Ni3Al calculated in the research coincide well with other theoretical and experimental values, and all the elastic constants of each interstitial phase satisfy the criteria for mechanical stability. Adding interstitial boron and carbon can reduce the various elastic moduli of Ni but the reduction induced by boron is more significant than that caused by carbon.
Boron and carbon have different effects on the elastic modulus of Ni3Al. Unlike interstitial carbon, in which the various elastic moduli of Ni3Al are enhanced, interstitial boron exerts a reduction effect on all the elastic moduli of Ni3Al except the bulk modulus, which is slightly improved. By calculating the values of the G/B and Poisson’s ratio of each system, it is found that interstitial boron and carbon can reduce the brittleness of Ni and so increase its ductility. Boron also increases the ductility of Ni3Al, while carbon hardly has any effect on brittleness or ductility.
Due to the anisotropic nature of the crystal, interstitial boron and carbon have different effects on the shear moduli and Young’s moduli of Ni and Ni3Al in crystals with different orientations. However, both the elements weaken the anisotropies of Ni and Ni3Al.
In order to analyze the effects of interstitial boron and carbon on the elastic properties of Ni and Ni3Al more fundamentally, the PDOSs and differential charge densities of the interstitial phases are analyzed. A strong hybridization effect between the 2p orbitals of the interstitial atoms and the 3d orbitals of their NN Ni atoms is discovered. Furthermore, the strength of the hybridization between Ni and C is more intense than between Ni and B. Also, the charge distributions between the interstitial atoms and their NN Ni atoms are endowed with obvious directivities, indicating that there are covalent-like bonds. The Ni–C bonds appear to be more covalent than the Ni–B bonds. This explains why adding the boron and carbon can increase the phase stabilities of Ni and Ni3Al, and why the various elastic moduli of the interstitial phases doped with carbon are larger than those doped with boron. The effects of boron and carbon on the elastic properties of Ni and Ni3Al described in this work provide useful guidelines for developing new alloys in the future.
1 | |
2 | |
3 | |
4 | |
5 | |
6 | |
7 | |
8 | |
9 | |
10 | |
11 | |
12 | |
13 | |
14 | |
15 | |
16 | |
17 | |
18 | |
19 | |
20 | |
21 | |
22 | |
23 | |
24 | |
25 | |
26 | |
27 | |
28 | |
29 | |
30 | |
31 | |
32 | |
33 | |
34 | |
35 | |
36 | |
37 | |
38 | |
39 | |
40 | |
41 | |
42 | |
43 | |
44 | |
45 | |
46 | |
47 | |
48 | |
49 | |
50 | |
51 | |
52 | |
53 | |
54 | |
55 | |
56 | |
57 | |
58 | |
59 | |
60 | |
61 | |
62 | |
63 | |
64 | |
65 | |
66 | |
67 | |
68 | |
69 | |
70 | |
71 | |
72 |