1. Introduction
In recent decades, with the increasing number of spacecraft launched by various countries, the space environment has become increasingly complex, and the development of high-performance structural materials suitable for the space environment has become an inevitable choice for the spacecraft in the future [
1,
2,
3,
4]. For the critical components and structural parts of spacecraft, high-strength steel with higher density is the main material commonly utilized in the past, but the application of traditional steel is greatly restricted because of the demand for lightweight spacecraft. Therefore, there is an urgent need to develop high-strength and lightweight materials suitable for spacecraft structural parts. Many studies have shown that titanium alloys are expected to replace high-strength steel as critical components and structural materials in spacecraft, all because of their low density, high specific strength, good corrosion resistance, and low temperature performance [
5,
6,
7]. Among hundreds of titanium alloys, Ti
90Al
6V
4 (wt.%) is widely used owing to its excellent mechanical performance. However, the service life of Ti
90Al
6V
4 (wt.%) will be greatly reduced because of poor impact resistance in the face of space debris, narrow working parameter range, complex manufacturing processes, as well as large deformation, and high strain rate at high temperature, if it is applied to spacecraft structural parts [
8,
9,
10]. As is known to all that adding different elements to the alloys can greatly affect the lattice stabilities and mechanical properties of the alloys [
11,
12,
13,
14], therefore, it is expected to enhance the applicability of Ti
90Al
6V
4 (wt.%) in space by adding alloying elements. In 2012, from the perspective of adding zirconium element to Ti
90Al
6V
4 (wt.%), Jing et al. developed a new type of alloy—Zr
47Ti
45Al
5V
3 (wt.%) [
15]. The chemical properties of Zr and Ti are similar because these elements belong to the same primary group (IVB), have two kinds of crystal structure (α- and β-type), and can form an infinitely miscible solid solution [
16]. Compared with Ti
90Al
6V
4 (wt.%), Zr
47Ti
45Al
5V
3 (wt.%) has great mechanical properties. The hardness of Zr
47Ti
45Al
5V
3 (wt.%) is above 400 HV, which is nearly 30% higher than Ti
90Al
6V
4 (wt.%). Its material strength after heat treatment is above 1400 MPa, and its density is only about 2/3 of high-strength steel [
17], thus, Zr
47Ti
45Al
5V
3 (wt.%) is a new type of alloy with great potential.
Since the Zr
47Ti
45Al
5V
3 (wt.%) was proposed, there are research findings which reported the relation of macroscopic rheological characteristic with microstructures of Zr
47Ti
45Al
5V
3 (wt.%) subjected to hot tensile, compression, or torsion deformation. In 2012, Liang et al. [
16] studied the structural and mechanical properties of hot-rolled Zr-Ti-Al-V alloys based on X-ray diffraction and metallographic analyses. Results showed that the microstructure of alloys went through a series of changes from the α (α’)-phase to part or all of the β-phase, thus affecting the mechanical properties as Zr increased. In 2013, Liang et al. [
18] experimentally studied the effect of Al content on the microstructure and mechanical properties of hot rolled Zr
47Ti
45Al
5V
3 (wt.%), results showed that the Zr-Ti-Al-V alloys with Al contents between 3.3 wt.% and 5.6 wt.% have excellent mechanical properties. Zhang et al. [
19] conducted impact compression experiments on Zr
47Ti
45Al
5V
3 (wt.%), and results showed that no obvious evidence of phase transition is found in the shock compression pressure range. In 2016, Tan et al. [
20] studied the hot deformation behavior of Zr
47Ti
45Al
5V
3 (wt.%) with an initial lamellar α-phase structure through a compression test in the temperature range of 823 K to 1073 K, researches showed that the flow curves exhibited a continuous flow softening in the α + β phase field. Liang et al. [
21] investigated the effects of V content (0–7 wt.%) on the structure and mechanical properties of Zr-Ti-Al-V alloys, and results showed that the phase composition changes as α’→α’+ (α’’ + β)→α’’ + β→β+ (α’’)→β with increased V content from 0 wt.% to 7 wt.%, and the elastic modulus and yield strength show an antiparabolic trend with increased V content. In 2018, Tan et al. [
22] studied the effect of thermal deformation on the α→β phase transition of Zr
47Ti
45Al
5V
3 (wt.%), results showed that with decreasing strain rate and increasing deformation temperature, the volume fraction and size of globular α phase increased.
At present, most of the research [
23,
24,
25] on experimental methods discussed the influence of alloy processing technology on structure and mechanical properties. For Zr-Ti-Al-V alloys, there is a lack of in-depth research about the effect of element composition and phase structure on their mechanical properties. However, mastering the relation between the element composition, phase structure, and mechanical properties is an effective way to optimize the alloy properties [
26,
27,
28,
29,
30,
31,
32], and it is of great significance to study the lattice stabilities and elastic properties of Zr-Ti-Al-V alloy systems. Marker et al. [
33] studied the elastic properties of the Zr-Ti-X (X = Nb, Sn, Ta) alloys in the BCC structure, and predicted the single crystal elastic stiffness coefficients and polycrystalline aggregate properties using DFT-based first-principles calculations. Liao et al. [
34] established the single-crystal elastic constant model of the quaternary BCC Zr-Ti-Nb-V alloys across full composition space using first principles, based on the research of Marker et al. [
33], the accuracy of the model was confirmed by comparing with previous calculations and experiments. Konopatsky et al. [
35] carried out the multi-cycle mechanical testing on ternary Ti
68Zr
18Nb
14 (at.%), quaternary Ti
68Zr
18Nb
13Ta
1 (at.%), Ti
68Zr
18Nb
12Ta
2 (at.%), and Ti
68Zr
18Nb
11Ta
3 (at.%) alloys, during multi-cycle mechanical testing, the following regularity appeared for all the alloys tested: during 10–15 first loading-unloading cycles, the mechanisms of plastic deformation and stress-induced martensite formation and stabilization prevail.
Many researchers have proved that phase composition of the Zr-Ti-Al-V alloys significantly depends on the amount of Zr and Ti [
16], and changes of phase compositions and structure with alloying element distinctly affected mechanical properties of Zr-Ti-Al-V alloys [
21]. It is of great significance to study the alloying effects of Al and V to provide guidance for further optimizing the lattice stabilities and mechanical properties regulation of quaternary Zr
47Ti
45Al
5V
3 (wt.%), in other words, the research of alloying element content could provide direction and assistance in guiding the research of quaternary alloys from the point of view of ternary alloys.
In this paper, we performed first-principles studies of Zr-Ti-X (X = Al, V) alloys using the special quasi-random structure (SQS) models to explore the relationship between element composition, phase structure, lattice stabilities, and elastic properties based on previous studies [
36]. This article starts from the basic composition of Zr
47Ti
45Al
5V
3 (wt.%), the lattice constants, total energies, and elastic constants of the Zr-Ti-X (X = Al, V) alloys are calculated. These calculations provide an effective guide for further optimizing the composition of the Zr-Ti-X (X = Al, V) alloys, and provide guidance and help in exploring the complex relationship between the ternary alloys and quaternary alloys.
2. Materials and Methods
Based on the ratio of the Zr atomic fraction and the Ti atomic fraction in the Zr47Ti45Al5V3 (wt.%), in order to investigate the effect of alloying elements Al and V on the crystal structures and mechanical properties of Zr47Ti45Al5V3 (wt.%) alloy, two types of ternary alloy systems were designed—Zr-Ti-Al alloy systems and Zr-Ti-V alloy systems. For these two types of alloy systems, the hcp and bcc structures were studied to model α and β phase because a mixture of two phases usually appears in these alloys, which gives a better understanding of the phase properties of two competitive phases. In this paper, the special quasi-random structure (SQS) method was used in the alloy systems with hexagonal close-packed structure (HCP, α phase) and body-centered cubic structure (BCC, β phase).
To simplify the calculation, the atomic ratio of Zr and Ti is approximately 0.35:0.50.
Figure 1 shows the SQS model of the Zr
35Ti
50Al
15 (at.%) alloys in HCP structure and BCC structure used in modeling in this article; the SQS model of the Zr
35Ti
50V
15 (at.%) alloys in HCP structure and BCC structure is shown in
Figure 2.
Table 1 shows the
(at.%) alloys (y:z = 0.35:0.50, and the atomic fraction of Al goes from 0 to 15 at.% with a gradient of 3 at.%) in HCP and BCC structures.
Table 2 shows the
alloys (at.%) (y:z = 0.35:0.50, and the atomic fraction of V goes from 0 to 15 at.% with a gradient of 3 at.%) in HCP and BCC structures. In this paper, the randomness of atomic configuration was reproduced while considering the local relaxation in the supercell through modeling.
In this article, the CASTEP (Cambridge Sequential Total Energy Package, Cambridge, UK) total energy calculation software package, based on the density functional theory (DFT) [
37] pseudopotential plane wave method [
38], was used to analyze the Zr-Ti-Al alloys and Zr-Ti-V alloys in HCP and BCC structures for first-principles calculations. Density functional theory (DFT) as an effective approximation method is used to solve the structure of multi-electronic systems [
39,
40,
41,
42,
43,
44,
45,
46], and it obtains the total energy of multi-electronic systems based on the energy minimization method. The pseudopotentials describe the valence configurations of the elements, which are as follows: Zr 4s
24p
64d
25s
2, Ti 3s
23p
63d
24s
2, Al 3s
23p
1and V 3s
23p
63d
34s
2. The electronic exchange-correlation energy was treated within the generalized gradient approximation (GGA) with the PBE revised for solids [
47]. The convergence criteria of 5 × 10
−7 eV per atom was applied to be self-consistent, all calculations were executed in reciprocal space to improve the speed and accuracy, in addition, the calculation and convergence were controlled by the cut-off point of the wave energy of the Ultrasoft pseudopotential plane. After convergence is tested, the Monkhorst-Pack grid of 2 × 2 × 3 is adopted for the brillouin zone integration, the plane wave kinetic energy cut-off was determined as 350 eV, and the unit cell parameters of Zr-Ti-Al alloys and Zr-Ti-V alloys are calculated by first principles. About the lattice stability, formation energy
was defined by:
Here, is the total energy per atom of the (at.%) alloy with alloying element concentration. , and are the energies per atom of Zr, Ti, and X in their ground state structures.
In this paper, according to our modeling system based on static modeling, the total energy and forming energy of the system calculated in this paper are considered from the level of electron exchange. They are themselves relative values because most of the energy is subtracted based on mass–energy equivalence. The energy of the atom itself is approximated as the absolute value of the total non-electron energy, which may not be considered in this article. We performed a static total energy calculation, based on that, the lattice stabilities of Zr-Ti-X (X = Al, V) alloys are evaluated with analyzing the total energy dependence on elements species, alloying fraction, and phase structure using the first principles in our work.
We determined the independent elastic coefficients, C
ij, based on expanding the total energy, U, of the crystal as a function of lattice strain Є [
48], which is expressed as
where V is the volume and U
0 is the total energy of the undistorted lattice system. The linear term in the expansion vanished, and the cubic or higher powers of ε
i terms in the polynomial expansion are neglected for sufficiently small strains.
In this paper, the elastic constant, C
ij, that describes the response of the crystal to external forces, is generally used to determine the strength of a material [
49]. Based on the research findings of Zhou et al. [
50], we theoretically predicted the elastic properties of Zr-Ti-X (X = Al, V) alloys of the alloying elements.
A small strain (δ) is applied to the solid lattice of the
(at.%) alloys to construct a quadratic function relationship between the total energy changes (∆U) and the strains (δ) according to the Hooks law, as follows:
where ∆U is the total energy changes, V is the original unit cell volume, C
ij are the elastic constants, ε
i and ε
j are the components of strain matrix.
In order to investigate these elastic constants, as shown in
Table 3, three sets of specific strains (δ) are applied to BCC metal in different directions. We also list the different strain configurations to compute the five independent elastic constants of HCP metal in
Table 4. The elastic constants can be obtained by fitting the relationship between the total energy changes (∆U) and the applied strains (δ).
For different unit cell structures, the stability criteria are also different according to Born-Huang’s mechanical stability theory. In this paper, the HCP Zr-Ti-Al alloys and HCP Zr-Ti-V alloys are calculated by first principles to obtain five independent elastic constants, C
11, C
12, C
13, C
33, and C
44, their stability is determined from the elastic constants by:
The BCC Zr-Ti-Al alloy systems and BCC Zr-Ti-V alloy systems are calculated by first principles to obtain three independent elastic constants, C
11, C
12 and C
44, their stability is determined from the elastic constants by:
The mechanical parameters of the Zr-Ti-Al alloys and Zr-Ti-V alloys can be calculated by using single crystal elastic constants according to the Voigt-Reuss-Hill scheme [
51,
52,
53,
54]. The bulk modulus (B), shear modulus (G), Young’s modulus (E), and Poisson’s ratio (ν) of the Zr-Ti-Al alloys and Zr-Ti-V alloys are calculated as follows:
The ratio of bulk modulus to shear modulus (B
H/G
H) proposed by Pugh [
55] has been widely used to evaluate the toughness and brittleness of materials. It should be noted that a lower value of B means the atomic bond is weaker. Consequently, materials with a low value of B exhibit relatively poor resistance to various forms of localized corrosion (such as intergranular corrosion, pitting corrosion and stress-corrosion cracking) [
56,
57,
58]. Furthermore, Clerc et al. [
59] have demonstrated that the hardness of the annealed metal is proportional to G. It is commonly known that materials with a high value of the B/G ratio (>1.75) often result in ductility, while materials with a low value of the B/G ratio (<1.75) often exhibit brittleness [
56,
58,
60,
61]. Vitos et al. investigated the elastic properties of the
alloys (13.5 at.% < c < 25.5 at.% and 8 at.% < n < 24 at.%) [
56], in that work, shear modulus (G) and bulk modulus (B) fall into the range of 74–81 GPa and 161–178 GPa, respectively.
In addition, the brittleness and ductility of materials can also be judged according to the value of Poisson’s ratio ν and the Cauchy pressure, that is, the difference between C
12 and C
44 [
62]. The larger the value of Poisson’s ratio ν, the better the ductility of the material, and if the calculated Cauchy pressure of materials is negative or their Poisson’s ratio ν is lower than 0.26, these materials can be considered brittle. Jiang et al. [
63] investigated the mechanical properties and Debye temperature of W-TM (TM = Cr, Cu, Fe, Mn, Mo, and Ni, respectively) alloys based on the first principles method, and results showed that the ductile/brittle properties of W-TM alloys can be evaluated based on the mechanical characteristic of B/G ratio, Poisson’s ratio (ν), and Cauchy pressure (C’).