† Corresponding author. E-mail:
Project supported by the National Key Research and Development Program of China (Grant No. 2017YFB0701502).
NiCo-based superalloys exhibit higher strength and creep resistance over conventional superalloys. Compositional effects on elastic properties of the γ and γ′ phases in newly-developed NiCo-based superalloys were investigated by first-principles calculation combined with special quasi-random structures. The lattice constant, bulk modulus, and elastic constants vary linearly with the Co concentration in the NiCo solution. In the selected (Ni, Co)3(Al, W) and (Ni, Co)3(Al, Ti) model γ′ phase, the lattice constant, and bulk modulus show a linear trend with alloying element concentrations. The addition of Co, Ti, and W can regulate lattice mismatch and increase the bulk modulus, simultaneously. W-addition shows excellent performance in strengthening the elastic properties in the γ′ phase. Systems become unstable with higher W and Ni contents, e.g., (Ni0.75Co0.25)3(Al0.25 W0.75), and become brittle with higher W and Co addition, e.g., Co3(Al0.25 W0.75). Furthermore, Co, Ti, and W can increase the elastic constants on the whole, and such high elastic constants always correspond to a high elastic modulus. The anisotropy index always corresponds to the nature of Young’s modulus in a specific direction.
Hardness and ductility are two of the key parameters for the design and characterization of materials and these quantities largely depend on the elastic properties. New NiCo-based superalloys[1–4] show superior mechanical properties and creep properties: other researchers investigated the Ni–Co–Al–X quaternary systems and their distribution coefficients, using experimental investigation and thermodynamic modelling of Ni–Co–Al–W[5] and Ni–Al–Ti.[6] Experiments on Ni–Co–Al–W–Cr quinary model superalloys[5,7] have also been reported. An increase of Co-addition in Ni–Co–Al–W systems decreases the γ′ solvus temperature and leads to a higher γ′ volume fraction.[7] Ti-addition increases the anti-phase boundary energy of the γ′ phase while reducing the alloy density,[8] but an excessive amount of Ti can form a D024 Ni3Ti (η) phase. The disordered face-centred cubic (fcc) γ and L12 γ′ phases formed with an fcc/L12 structure formed in the Ni–Al–Co–Ti quaternary alloy at 750 °C to 1100 °C in tests.[9] Atomic probe tomography shows that Co addition (> 19%) decreases the solubility of Al and Ti in the γ′ phase and caused Ni3(Ti, Al) transformation in (Ni, Co)3(Al, Ti) systems.[10] However, reports of concentration effect on mechanical properties of NiCo-based superalloys by first principles calculation remain sparse.
In the γ′-Co3(Al, W) phase[11] of Co-based superalloys, Al and W randomly occupy the Al-sublattice. In NiCo-based superalloys, the precipitation of Ni and Co is complicated, they both randomly occupy sites in the γ phase and also exist in caused γ′ phases. The first problem when studying NiCo-based superalloys is how to deal with the random problem caused by first principles calculation: the presence of a configurational substitutional disorder leads to a loss of translational periodicity. The prediction of the elastic properties of low-symmetry systems, such as random alloys, is less straightforward.
The widely used techniques for modelling disordered alloys are special quasi-random structures (SQS),[12] coherent potential approximation (CPA),[13] cluster expansion (CE),[14] and virtual crystal approximation (VCA).[15] The random distribution of impurity atoms in the CPA is elucidated in the framework of the mean-field approximation and ignore local lattice relations. The CE is based on the assumption that the energy (of any other scalar property) of the system can be expanded in terms of a set of well-chosen structural motives (figures) and the expansion parameters are typically obtained by fitting the CE Hamiltonian from ab initio electronic structure calculations. In the SQS, the randomness is introduced by mimicking as closely as possible the correlation functions of an infinite random alloy within a finite supercell. The calculation of electronic structure in SQS also involves local relaxation effects.
Here, we use first-principles calculation combined with SQS to investigate the strengthening mechanisms related to alloying elements and concentration on the elastic properties of NiCo-based superalloys, including the γ -NiCo phase and γ′-(Ni, Co)3(Al, W/Ti) phase. The paper is organised as follows: Section
Special quasi-random structure SQS[12,16] is a “special” N-atom periodic quasi-random structure S whose correlation function
The elastic constants can be calculated from the stress–strain, or strain–energy, methods. Since the total energy depends on the volume much more strongly than that on the strain,[18] we calculated elastic constants from the energy expansion using volume conserving deformations, i.e., monoclinic (C44) and orthorhombic distortions (C11–C12).
The 21 inequivalent elastic constants of a triclinic system can be given as a vector with 21 components. In general, projector Psym gives the closest elastic tensor with higher point group symmetry Xsym as
A series of elastic properties can be calculated by the elastic constants, i.e., directional Young’s modulus, shear modulus in a specific direction, and the elastic modulus of the polycrystalline materials.
The Young’s modulus is the ratio of uniaxial stress to stain measured along the same axis and can be expressed as[22,23]
The shear modulus G along [110] and [11
For a polycrystalline aggregate, the principal crystallographic axes are mostly randomly oriented in space and the material is statistically isotropic. The bulk modulus B, shear modulus G, and anisotropy index AZ are determined from elastic constants by Voigt–Reuss–Hill approximation.[22,27–29]
Here, we selected the possible systems in NiCo-based superalloys, including: (i) NiCo solution (γ phase) both in the paramagnetic state and ferromagnetic state; (ii) binary L12 systems (γ′ phase), including Co3Al, Ni3Al, Co3 W, Ni3 W, Co3Ti, and Ni3Ti; (iii) ternary L12 systems (γ′ phase), including (Ni, Co)3Al, Co3(Al, W), Ni3(Al, W), Co3(Al, Ti), and Ni3(Al, Ti); and (iv) quaternary L12 systems (γ′ phase), including (Ni, Co)3(Al, W), and (Ni, Co)3(Al, Ti).
In a real alloy, the γ phase is an fcc random solution with Ni and Co occupying sites randomly when the Co concentration is below 65%. The shape of the (i) system supercell is 2[001] × 2[010] × 2[100], including 32 atoms as shown in Fig.
The MedeA software give access to the 32-atom SQS in the whole concentration for the calculation of (i) systems (Fig.
After relaxation of the SQSs, the equilibrium lattice constants (a) and bulk modulus (B) were obtained by calculating the energy–volume dependence of the system and fitting it to the Murnaghan equation of state.[34] In the calculations of the elastic constants with cubic symmetry, a strain–energy approach is applied. Six types of deformations were adopted in the calculation of the elastic constants.
To obtain the total energies and extract the elastic constants of the supercells, density functional theory (DFT) calculations are conducted[35] by using the Vienna Ab initio Simulation Package (VASP),[36,37] a projector augmented wave method,[38] and the generalized gradient approximation (GGA)[39] in the parameterization introduced by Perdew, Burke, and Ernzerhop (PBE). The minimum cut-off of the plane wave energy is 350 eV. The spacing between k-points is 0.18 Å−1. The spin-polarized calculation is considered in NiCo solution, in which the initial magnetic moment of Ni and Co are 0.65 μB and 1.6 μB, respectively.
The calculated results are in good agreement with previous reported results (see Table
From Fig.
From the definition of the lattice mismatch δ, we obtain,
The addition of Co, Ti, and W all causes to an increase in bulk modulus (Fig.
In Fig.
The bulk modulus of free-electron-like metals varies significantly with the atomic volume V when different elements are compared, and V does not depend much on the lattice structure as long as the electronic structure is not much changed.[22] We tested the relationship utilizing the bond valence based on the uniform electron gas model for transition metals.[49] Based on a published calculation method[41] and electron density results,[50] figure
Before providing further analysis of the elastic properties, we analyze their stability and ductility.
In the Ni–Co binary phase diagrams, the NiCo solution is an fcc structure with a Co concentration below 65%. In the binary systems, the stable structure of Ni3Al and Co3Ti is L12, the stable structure of Co3W is D019, the stable structure of Ni3Ti is D024, and L12 is not a stable structure for Co3Al. In the Ni–W binary phase diagrams, there are stable intermetallic compounds of Ni4W and Ni2W present, and the stable structure of Ni3W is D022.[51] Here, we provide a stability analysis of ternary and quaternary systems with L12 crystal structures, and do not do so for NiCo solution and binary systems. Three methods were adopted to measure the structural stability, i.e., the Born stability criteria,[52] formation energy, and pseudo-gap.
The Born stability criteria for cubic crystals are
The formation energy of the alloy was calculated using the constituent pure elements as reference states as follows:
Here, Pugh’s approximation[30] and the Cauchy pressure are adopted to evaluate the mechanical properties of the selected systems with various alloying concentrations,[54,55] as shown in Fig.
From the elastic constants results of Subsection
In Fig.
The {110} plane includes the cubic high symmetry crystallographic directions [100], [110], and [111]. Here, we chose this cross-section to give a complete presentation of the directional Young’s modulus. From the elastic constants results, we only give the Young’s modulus of NiCo(PM), NiCo(FM), Co3(Al, W), (Ni, Co)3W, Co3(Al, Ti), and (Ni, Co)3(Al, W) systems. As all these systems satisfy 2S11 − 2S12 − S44 > 0 in Fig.
Except for the inclusion of the main crystal direction, the {110} plane also includes tetrahedral and octahedral interstices of the cubic systems. Here, we chose the {110} plane to analyze the charge density difference (Fig.
In the bonding as shown in Fig.
The lattice constant, bulk modulus, and elastic constants of NiCo-based model systems were investigated by first principles calculation in combination with SQS. The relationship between bulk modulus and volume, stability and ductility, and elastic modulus were discussed. The key results may be summarised as follows.
[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] |