Next Article in Journal
The Effect of LaPO4 Crystal Morphology on Gas-Phase Catalytic Synthesis of Anisole
Previous Article in Journal
Remineralizing Effect of Three Fluorinated Varnishes on Dental Enamel Analyzed by Raman Spectroscopy, Roughness, and Hardness Surface
Previous Article in Special Issue
Influence of Cooling Methods on Microstructure and Mechanical Properties of TiB2@Ti/AlCoCrFeNi2.1 Eutectic High-Entropy Alloy Matrix Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Approach of Stability and Mechanical Properties in (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) Refractory High-Entropy Alloys

1
College of Materials Science and Engineering, Hunan University, Changsha 410082, China
2
Xinjiang Xianghe New Materials Technology Co., Ltd, Hami 839000, China
3
Research Institute of Automobile Parts Technology, Hunan Institute of Technology, Hengyang 421002, China
4
School of Material Science and Hydrogen Energy Engineering of Foshan University, Foshan 528001, China
*
Authors to whom correspondence should be addressed.
Coatings 2025, 15(9), 1092; https://doi.org/10.3390/coatings15091092
Submission received: 22 August 2025 / Revised: 11 September 2025 / Accepted: 15 September 2025 / Published: 17 September 2025
(This article belongs to the Special Issue Innovations, Applications and Advances of High-Entropy Alloy Coatings)

Abstract

The stability and mechanical properties of (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) refractory high-entropy alloys have been investigated by combining the first-principles with special quasi-random structure (SQS) method. It is found that with the increase in solute concentration x, the ΔHmix of (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) linearly decreases, whereas both ΔHmix and ΔSmix of (TiZrHf)1−x(NbTa)x increase initially and subsequently decrease, with the crossover occurring at x = 0.56. The ΔHmix of (TiZrHf)1−x(NbTa)x and (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) alloys are larger and lower than that of TiZrHf, respectively, while the ΔSmix of all (TiZrHf)1−x(AB)x is larger than that of TiZrHf. The formation possibility parameter Ω of all (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) first decreases sharply, followed by a gradual decrease. And the local lattice distortion (LLD) parameter δ remains relatively stable around x = 0.56 for all cases, after which it decreases sharply until x = 0.89. The δ value of (TiZrHf)1−x(AB)x is higher than that of TiZrHf for x < 0.56 but becomes lower beyond this composition. The valence electron concentration (VEC), a possible indicator for a single-phase solution, of (TiZrHf)1−x(AB)x increases nearly linearly, while the formation energy ΔHf of (TiZrHf)1−x(AB)x shows the opposite tendency, except for (TiZrHf)0.67(NbTa)0.33. Furthermore, the VEC of all (TiZrHf)1−x(AB)x alloys increases, whereas their ΔHf decreases compared to that of TiZrHf. The ideal strength σp of (TiZrHf)1−x(AB)x increases linearly, reaching approximately 2.12 GPa. The bulk modulus (B), elastic modulus (E), and shear modulus (G) also exhibit linear increases, and their values in all (TiZrHf)1−x(AB)x alloys are higher than those of TiZrHf, with some exceptions. The Cauchy pressure (C12C44) and Pugh’s ratio G/B of all (TiZrHf)1−x(AB)x alloys increase, whereas the Poisson’s ratio ν exhibits the opposite trend. Moreover, the C12C44 and G/B ratio of TiZrHf are lower and higher, respectively, than those of (TiZrHf)1−x(AB)x, and the ν of TiZrHf is lower than that of (TiZrHf)1−x(AB)x. This study provides valuable insights for the design of high-performance TiZrHf-based refractory high-entropy alloys.

Graphical Abstract

1. Introduction

Refractory high-entropy alloys (RHEAs) are a novel class of materials that have been widely studied due to their high yield strength and excellent high-temperature softening resistance, which significantly broadens the application range of traditional alloys [1,2], and possibly replace nickel-based superalloys as a candidate material [1,3,4]. Additionally, coating technology integrates the material advantages of RHEAs with practical application requirements. Through advanced surface engineering techniques such as laser cladding, plasma spraying, and vapor deposition, high-performance protective coatings can be fabricated on substrate surfaces, including superalloys, titanium alloys, and carbon-based materials. This approach effectively enhances the durability and service life of components operating in extreme environments, such as high-temperature, corrosive, or abrasive conditions, without significantly increasing production costs or compromising the intrinsic properties of the base materials [5,6]. Popular RHEAs with high mixing entropy ΔSmix > 1.5R (R = 8.314 J/(mol·K) [7,8] are usually composed of five refractory elements containing Nb, Mo, W, Ta, etc. to form single-phase solid solutions with simple face-centered cube (FCC) or body-centered cube (BCC) structures [1,9]. Although majority of RHEAs exhibit good strength and resistance to high temperatures, they are rather brittle at room temperature. For example, the yield stress of a high-entropy Nb20Mo20Ta20W20V20 alloy at room temperature can achieve 1246 MPa, while the ductility along the almost parallel compression direction is only 1.5% [10].
Fortunately, the design and development of TiZrHf-based RHEAs bring new hope to solve the brittleness of RHEAS as Hf effectively improves the ductility in the BCC system [11,12]. Dirras et al. [13] adopted vacuum arc melting technology to prepare a Ti20Zr20Hf20Nb20Ta20 high-entropy alloy in experiments, and its elastic and plastic behaviors are investigated by the pulse-echo technique and tensile experiments. Their results show that the bulk modulus (B), shear modulus (G), and Young’s modulus (E) of the Ti20Zr20Hf20Nb20Ta20 alloy are 134.6, 28.0, and 78.5 GPa, respectively. The results further showed that the Ti20Zr20Hf20Nb20Ta20 alloy has a low Pugh ratio (B/G) and positive Cauchy pressure (C12–C44) of 0.208 and 80 GPa, respectively, suggesting that the Ti20Zr20Hf20Nb20Ta20 alloy is a ductile material. The tensile test results show that the yield strength of the Ti20Zr20Hf20Nb20Ta20 alloy can reach 800~840 MPa. Sheikh et al. [12] designed a high-entropy Hf0.5Nb0.5Ta0.5Ti1.5Zr alloy with excellent strength-ductility by combining theory and experiment. They compared the valence electron concentration (VEC) of a range of medium- and high-entropy alloys, showing that reducing VEC can lead to a transition from brittle to plastic in multi-component alloys. Thus, they prepared Hf0.5Nb0.5Ta0.5Ti1.5Zr RHEAs with low electron concentration (4.25) with arc melting technology. Tensile experiments showed that the yield stress, fracture stress, and elongation of Hf0.5Nb0.5Ta0.5Ti1.5Zr RHEA are 903 MPa, 990 MPa, and 18.8%, respectively. Furthermore, achieving the correct chemical composition is an effective way to maintain the BCC structure of TiZrHf-based RHEAs, and avoid the formation of brittle ω phase with the hexagon close-packed (HCP) phase [14,15]. In addition, the stability of TiZrHf-based RHEAs is strongly affected by local lattice distortions (LLDs), especially for BCC structure [16]. Therefore, it is urgent to study novel TiZrHf-based alloys in a variety of chemical compositions for achieving better RHEAs and to establish the relevant database.
Generally speaking, investigating the mechanical properties of materials can be divided into two types: (1) small strain response in Hooke’s law; (2) nonlinear response stress at larger strains [17]. The small strain response mainly studies the elastic modulus of materials containing bulk, shear, and Young’s moduli, and the nonlinear response stress at larger strains called is ideal strength. However, because of the multi-variable nature and complexity of the experiment, accurately investigating the mechanical properties of RHEAs is quite difficult [18,19,20]. In the past decade, with the emergence of theoretical calculations, e.g., first-principles (FP) calculation, they have played a crucial role in multiple material fields [21,22,23,24,25,26,27,28,29].
Very recently, Li et al. [30] investigated the phase stability and the micromechanical properties of (TiZrHf)1−xTMx (TM = (V, Nb, Cr, Mo, W)) based on first-principles method combing coherent-potential approximation (CPA). Their results suggested that Mo and W have a stronger stability for BCC structure, and the elastic hardening effect of Mo is more pronounced than that of Nb. In the framework of density functional theory (DFT), Chen et al. [31] employed thermodynamic models to investigate the stability of the Hf-Nb-Ta-Ti-Zr HEA alloy with BCC structure. The results showed that the equiatomic HfNbTaTiZr HEA suffers from phase decomposition below critical temperature of 1298 K, and the HEA decomposes most favorably to BCC NbTa-rich and HfZr-rich phases. Meng et al. [32] employed the DFT calculation based on special quasi-random structure (SQSs) to study the influences of LLDs on elastic properties for both hexagonal TiZrHf(Sc) alloys. They discovered that the LLDs significantly reduce the shear elastic properties, and the influence of LLDs on elastic properties for TiZrHf is more significant than that for TiZrHfSc.
However, until now, a comprehensive study of the stability and mechanical properties in relation to the chemical composition of TiZrHf-based RHEAs has not been reported. Here, we selected three substituted elements (Nb, Mo, and Ta) to construct (TiZrHf)1−x(AB)x (x = NbMo, NbTa, MoTa) five-element RHEAs based the SQS method and investigated their stability and mechanical property by using the first-principles calculation. Meanwhile, we chose eight chemical components, containing x = 0.11, 0.22, 0.33, 0.44, 0.56, 0.67, 0.78, 0.89, for (TiZrHf)1−x(AB)x alloys, and calculated the formation energy Hf, entropy of mixing ΔSmix, enthalpy of mixing ΔHmix, parameter Ω, standard deviation δ-parameter, bulk modulus B, shear modulus G, Young’s modulus E, Cauchy pressure C12C44 and Poisson’s ratio ν, and ideal tensile strength σp as a function of concentration x to evaluate their stability and mechanical properties.

2. Calculation Details

In this work, the first-principles calculation was implemented in the Vienna Ab initio Simulation Package (VASP) [33] based on the density functional theory (DFT). Meanwhile, the plane-wave basis projector augmented-wave (PAW) method [34] and the Perdew–Burke–Ernzerhof (PBE) of generalized-gradient approximation (GGA) [35] were, respectively, applied to account for the core-valence electron interaction and exchange-correlation functional. During the whole relaxation of initial models, the cut-off energy of plane-wave was 350 eV, and the PAW_GGA pseudopotentials of (Ti-Zr)_sv and (Nb-Ta)_pv were adopted. Furthermore, we tested the total energy of (TiZrHf)0.44(AB)0.56 (AB = NbMo, NbTa) as a function of cutoff energy, and it is plotted in Figure 1a,b. The results indicate that when the cutoff energy exceeds 250 eV, the total energy exhibits minimal variation. Therefore, a cutoff energy of 350 eV was deemed sufficient and was employed in this study. For the accuracy of all calculations, the total energy and Hellmann–Feynman forces [36] on each atom were, respectively, converged to 10−6 eV/atom and 0.01 eV/Å by the conjugate gradient (CG) minimization method.
Here, we established the models of (TiZrHf)1−x(AB)x RHEAs with the BCC structure by fixing the proportional concentrations of Ti, Zr, and Hf to adding three combinations for A and B elements with proportional concentrations (AB = NbMo, NbTa, MoTa and x = 0.11, 0.22, 0.33, 0.44, 0.56, 0.67, 0.78, 0.89). Note: The proportional concentrations of the elements in the parentheses are equimolar. To solve the random arrangement and chemical disordering of (TiZrHf)1−x(AB)x alloys, we adopted the special quasi-random structure (SQS) method carried out in the alloy theoretic automated toolkit (ATAT) [37]. And the corresponding Brillouin zone integrations employed the Gamma centered mesh [38] of 4 × 3 × 3 k-point grid. Figure 2 shows all models of (TiZrHf)1−x(NbMo)x as examples. The strain matrix with a maximum strain of 9% was employed to calculate the elastic properties using the energy-strain method, while the ideal strength was determined under the condition of applying a maximum strain of no more than 35% along the [001] direction. All calculations were performed under ambient pressure conditions of 0 GPa.

3. Results and Discussion

3.1. Formation and Stability

To evaluate the possibility of formation of (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) RHEAs, the enthalpy of mixing ΔHmix firstly is calculated. Generally, when the ΔHmix is closer to 0 kJ/mol, single-phase solid solutions will be easier to form. The range of optical critical values of ΔHmix for forming (TiZrHf)1−x(AB)x is generally −15 kJ/mol < ΔHmix < 5 kJ/mol. The entropy of mixing ΔSmix is also a main factor, and the whole energy of system can be reduced by higher ΔSmix to form a simple solid solution. The ΔHmix and ΔSmix can be obtained by the following [20,39]:
Δ H m i x = i = 1 n 4 H i , j m i x C i C j
Δ S m i x = R i = 1 n C i l n C j
where Δ H i , j m i x are, respectively, the ith and jth mixing entropies from Miedema’s mode [40]. Ci and Cj are the atomic concentrations from ith and jth elements. R is the ideal gas constant. Note: All primary data for calculated Δ H i , j m i x can be obtained from ref [40].
To describe the possibility of (TiZrHf)1−x(AB)x to form a single solid solution structure (SSSS), we introduce an important formation parameter Ω that was proposed by Yang et al. [41], where a larger value of Ω indicates that a greater ΔSmix is to reduce more free energy, which makes the system more stable according to Gibbs free energy theory. It can be calculated as follows:
Ω = T m Δ S m i x Δ H m i x
T m = i = 1 n C i T e i
where Tm and Te are the melting temperatures of (TiZrHf)1−x(AB)x and the simple substance of the ith element from refs. [42,43].
In addition, for simple multi-component solid solutions, the effects of LLDs should be considered. A standard deviation δ-parameter of first-nearest neighbor distortions can characterize the LLDs. When ranges of Ω and δ are, respectively, greater than 1.1 and less than 6.6%. The simple solid solution phase is easier to form in multi-component alloys with the BCC system, and it can be calculated as follows [44]:
δ = 1 n n x n μ 2
where x and µ correspond to the nth data and the average value, respectively.
The formation enthalpy ΔHf can be calculated to describe the relative stability for different phases, whereas the average valence electron concentration (VEC) can also be used to evaluate the ability to form single-phase solutions. In the BCC system, it was found that BCC phases are stable at lower VEC (<6.87). They can be obtained and defined, as follows [45,46,47,48]:
Δ H f = E t o t a l n E i N
V E C = i = 1 n C i V E C i
where the Ei and n correspond to the energy of the ith element and the number of atoms of the ith element, respectively. Etotal and N are the total energy of systems and total number of atoms. The VECi is the VEC of each element i.
The formation energy of atomic vacancies E v a c plays an important role in computational materials science. It can directly or indirectly reflect the thermodynamic stability of materials and their resistance to phase transformation. In this work, the vacancy formation energy of (TiZrHf)1−x(AB)x is obtained by averaging the vacancy formation energies of all elements in the system, as follows:
E v a c = 1 i E i E p e r . + μ i N
where E i , E p e r . , N, and μ i , respectively, represent the total energy for the defective cell of the i-th element and perfect cell, the number of times of vacancies, and the chemical potential of the i-th element.
We have calculated the enthalpy of mixing ΔHmix, entropy of mixing ΔSmix, melting temperature Tm, SSSS formation possibility parameter Ω, LLD parameter δ, average valence electron concentration VEC, formation enthalpy ΔHf, and the formation energy of atomic vacancies Evac of (TiZrHf)1−x(AB)x RHEAs; the values of these parameters were summarized in Table 1. To clearly clarify the correlation between concentration x and the above parameters, Figure 3 shows the ΔHmix, ΔSmix, Ω, δ, VEC, and ΔHf as a function of concentration x. From Figure 3a, we can see that, with the increase in concentration x, the ΔHmix of (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) linearly decreases from the concentration x = 0.11 to 0.8 whereas the ΔHmix of (TiZrHf)1−x(NbTa)x firstly increases from the concentration x = 0.11 to 0.56, and then decreases to x = 0.89. It can be found that the ΔHmix of all (TiZrHf)1−x(NbTa)x alloys is larger than that of TiZrHf of 0 kJ/mol while the ΔHmix of all (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) is lower than that of TiZrHf of 0 kJ/mol. Figure 3b shows only a curve about ΔSmix as a function of concentration because the ΔSmix does not depend on the type of element. We can see that the calculated ΔSmix of (TiZrHf)1−x(AB)x, with the concentration increasing first, increases from the concentration of x = 0.11 to 0.56 and then decreases to 0.89. And the ΔSmix of all (TiZrHf)1−x(AB)x is larger than that of TiZrHf, except for (TiZrHf)0.11(NbMo)0.89 of 9.04 J/(mol·K).
According to Equations (1)–(3), we have further obtained the key parameter Ω based on the above results and drawn the relationship between the concentration and Ω in Figure 3c. From the results, the Ω of (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) first decreases sharply from x = 0.11 to 0.44 and then slowly decreases to 0.89, whereas the Ω of (TiZrHf)1−x(NbTa)x first decreases slowly from x = 0.11 to 0.56 and then increases to 0.89. However, the Ω of TiZrHf cannot be calculated as a ΔHmix of 0 kJ/mol forces when Ω approaches . The δ is also an important parameter based on Zhang et al.’s extensive research on multi-components with simple BCC, FCC, and HCP crystals.
In Figure 3d, we draw the curves of δ for (TiZrHf)1−x(AB)x alloys with the substituted concentration increasing. From the results, it can be seen that the δ of (TiZrHf)1−x(AB)x floats within the 0.56 concentration. After 0.56 concentration, the δ decreases sharply to 0.89. The δ of (TiZrHf)1−x(AB)x is larger than that of TiZrHf of 11.67% before x = 0.56. After that, the δ of cases is lower than that of TiZrHf of 11.67%, indicating that a higher substituted concentration may maintain the BCC structure [16]. It is known that when the average valence electron concentration (VEC) is less than 6.87 in BCC system, the structure is easier to form. But the relationship between stability and VEC for (TiZrHf)1−x(AB)x alloys has not been proposed. The calculated VEC and formation enthalpy Hf with the concentration increasing are plotted in Figure 3e,f. It can be found that the VEC increases linearly from x = 0.11 to 0.89 while the Hf shows the opposite, except for (TiZrHf)0.67(NbTa)0.33 of 7.71 kJ/mol. The VEC of TiZrHf is lower than that of all (TiZrHf)1−x(AB)x alloys whereas the ΔHf of TiZrHf is higher than that of all cases, except for (TiZrHf)1−x (NbTa)x (x = 0.11, 0.22, 0.33, 0.44). These results show that the lower VEC has a higher stability for (TiZrHf)1−x(AB)x RHEAs in the range of VEC. Further, the Evac of (TiZrHf)1−x(AB)x alloys is at the concentration of x = 0.56, drawn in Figure 3g. Obviously, the Evac of ~2.38 for (TiZrHf)1−x(AB)x is much larger than that of −0.24 for TiZrHf, and the (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) shows a larger Evac.

3.2. Ideal Strength

The ideal strength σp presents the upper strength limit along a fixed crystal orientation and provides insights into how the crystal is dissociated. We have calculated the σp by plotting stress–strain curves, given by the following:
σ = 1 V ε d E d ε
where ε, σ and V are the engineering strain, corresponding tensile stress and equilibrium volume of the supercell, respectively.
In Figure 4a–c, the stress–strain curves of (TiZrHf)1−x(AB)x are given by using Equation (9). We can find that, with the respectively increasing strains to 0.20 and 0.25 along the [001] direction, the σ of (TiZrHf)(NbMo)x (x = 0.11, 0.67, 0.78, 0.89), (TiZrHf)1−x(NbTa)x (x = 0.11, 0.33), (TiZrHf)1−x(MoTa)x (x = 0.11, 0.22, 0.33, 0.78) and (TiZrHf)1−x(NbMo)x (x = 0.22, 0.33, 0.44, 0.56), (TiZrHf)1−x(NbTa)x (x = 0.44, 0.56, 0.67, 0.78, 0.89), and (TiZrHf)1−x(MoTa)x (x = 0.56, 0.67, 0.89) gradually increases to the maximum value σp, and then dramatically decreases at the 0.25 and 0.30 strains, whereas the maximum strain for (TiZrHf)0.78 (NbTa)0.22 and (TiZrHf)0.56 (MoTa)0.44 corresponds to the σp when it is at 0.30. To describe the ideal strength σp of (TiZrHf)1−x(AB)x as a function of concentration x, Figure 4d shows the calculated σp with the concentration x increasing. We can clearly see that the σp of (TiZrHf)1−x(AB)x increases linearly in a range of ~ 2.12 GPa, suggesting that the higher the concentration x is, the more advantageous it is for (TiZrHf)1−x(AB)x deformation. Moreover, at all concentrations, the additions of NbMo, NbTa, and MoTa element combinations would increase the σp of TiZrHf.
To study the effect of stress on the structure and thermodynamics of materials, we use the Debye model implemented in Gibbs2 codes [49] to calculate key thermodynamic parameters, including the heat capacity CV, entropy S, thermal expansion coefficient αT and Gibbs free energy Eg, as a function of strains at the NbMo, NbTa and MoTa concentration x = 0.56 under room temperature. In Figure 5a–d, the αT of (TiZrHf)1−x(AB)x containing pure TiZrHf presents slight changes in varying degrees with the increase in strains while the CV, S, and Eg of all cases increases. Moreover, adding NbMo, NbTa, and MoTa would significantly decrease all thermodynamic parameters of TiZrHf. The (TiZrHf)1−x(MoTa)x and (TiZrHf)1−x(NbMo)x show, respectively, the lowest αT, Eg and CV, S.

3.3. Elastic Property

To study the elastic property of HEAs, we first calculate their elastic constants using the energy–strain method, and other mechanical parameters in this work are dependent on calculated elastic constants. Elastic constants are a measure of a material’s resistance to small deformations under a constant stress, and they can be calculated as follows:
Δ E = V 0 2 i , j = 1 6 C i j ε i ε j
where V0, Cij, and Δ E are the volume of the equilibrium cell, elastic constants, and total energy difference between the deformed cell and the perfect cell, respectively.
It is well known that the number of independent elastic constant components depends on the symmetry of the crystal. In a cubic crystal system, Cij contains three independents, C11, C12, and C44. And this Cij can be interpreted as an important judgment of material stability for (TiZrHf)1−x(AB)x alloys in the Born criterion, as follows:
C 11 > 0 , C 44 > 0 , C 11 C 12 > 0 , C 11 + 2 C 12 > 0
All values of Cij can be found in Table 2, and the calculated results show that all the (TiZrHf)1−x(AB)x phases containing pure TiZrHf can meet the mechanical stability. The C11 of TiZrHf is larger than that of (TiZrHf)1−x(NbTa, MoTa)x (AB = NbTa, MoTa; x = 0.11, 0.22) and (TiZrHf)1−x(NbTa)x (x = 0.11, 0.22, 0.33), and lower than that of the rest of (TiZrHf)1−x(AB)x. The C12 of all (TiZrHf)1−x(AB)x is larger than that of TiZrHf, and the C44 of TiZrHf is larger than that of (TiZrHf)1−x(NbTa, MoTa)x (AB = NbTa, MoTa; x = 0.11, 0.22, 0.33, 0.44, 0.56) and (TiZrHf)0.89(NbTa)0.11, and lower than that of the rest of (TiZrHf)1−x(AB)x. Furthermore, the bulk modulus B and shear modulus G have been calculated by the above Cij, given as follows:
B = B R = B V C 11 + 2 C 12 3
G V = C 11 C 12 + 3 C 44 5
G R = 5 × ( C 11 C 12 ) × C 44 4 × C 44 + 3 × ( C 11 C 12 ) ;
G = G V + G R 2
where   B R ,   G R and B V , G V are the lower and upper limits of polycrystalline elastic modulus and come from the Reuss and Voigt methods, respectively.
As previously mentioned, B and G are used to describe the measurement of resistance to volume and shear deformations under loading strains [50]. In addition, the Young’s modulus E and Poisson’s ratio ν are also important, and the E and ν can, respectively, reflect the forward deformation and heterogeneity of materials. They can be calculated by the following:
E = 9 G B 3 B + G
ν = 3 B 2 G 2 × ( 3 B + G )
Additionally, the Debye temperature θ D is also an important parameter that can be directly calculated from elastic constants. It enables the evaluation of the bonding strength between atoms in a material and provides data support for analyzing the material’s hardness, melting point, and thermal conductivity. It can be calculated as follows [51,52,53]:
θ D = h κ B 3 4 π 1 3 v m
v m = 1 3 1 v L 3 + 2 v S 3 1 3
v L = 3 B + 4 G 3 ρ 1 2
v S = G ρ 1 2
where h , κ B , v L , v S , and ρ are the Plank’s values, Boltzmann’s constants, longitudinal values, traverse sound velocities, and corresponding density of the crystal, respectively.
We have obtained the Cauchy pressure C12C44, bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Pugh’s G/B, Poisson’s ratio ν, and Debye temperature θ D , collected in Table 2 and Table 3 according to Equations (12)–(17) and using the results of Cij. To clarify the relation between mechanical behavior of (TiZrHf)1−x(AB)x and substituted concentration x, we have to first plot the bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa) as a function of x, respectively, as seen in Figure 6a. It can be seen that the B, E, and G of (TiZrHf)1−x(AB)x increase linearly with the increase in concentration x, and the B, E, and G of all (TiZrHf)1−x(AB)x are larger than that of TiZrHf, except for (TiZrHf)1−x(AB)x (x = 0.11) of B, (TiZrHf)1−x(NbTa)x (x = 0.11, 0.22, 0.33, 0.44, 0.56) of G and E, (TiZrHf)1−x(AB)x (AB = NbMo, MoTa; x = 0.11, 0.22) of G and E, and (TiZrHf)0.67(NbMo)0.33 of G. The results further show that when the element containing Mo is added, the elastic modulus is significantly improved.
To evaluate the atomic bonding characteristics and ductile behavior of (TiZrHf)1−x(AB)x, we have drawn the curves of the Cauchy pressure C12C44 and Pugh’s G/B as a function of concentration x in Figure 6b,c. Research showed that when C12C44 is greater than 0 GPa, the atomic bonding of the material exhibits a metallic characteristic; otherwise, it presents as a directional covalent characteristic [54,55]. And the larger absolute value indicates the stronger metallic or directional covalent characteristic of materials, whereas the larger G/B shows the more brittle. From the results, the C12C44 of all (TiZrHf)1−x(AB)x increases with the substituted concentration x, except for (TiZrHf)0.11(NbMo)0.89 of 84.16 GPa and (TiZrHf)0.78(NbTa)0.22 of 64.49 GPa and are greater than 0 GPa. Similarly, the G/B of all (TiZrHf)1−x(AB)x increases with the increasing concentration x, except for (TiZrHf)0.67(NbMo)0.33 of 0.308, (TiZrHf)0.56(NbMo)0.44 of 0.297, and (TiZrHf)0.67(NbTa)0.33 of 0.234. This phenomenon indicates that atomic bonds of (TiZrHf)1−x(AB)x present stronger metallic characteristics with the increase in concentration x, and the ductility improves. Further, the C12C44 and G/B of TiZrHf are lower and larger than that of (TiZrHf)1−x(AB)x, indicating that the atomic bonds of TiZrHf have a weaker metallic characteristic and it presents worse ductility. Poisson’s ratio ν is widely used to show the heterogeneity of materials [56], and when the ν is greater/smaller than 0.26, it shows uniform/nonuniform behaviors in response to applied stress. Figure 6d shows the calculated ν as a function of concentration x, and the ν of (TiZrHf)1−x(AB)x shows the opposite trend. Clearly, all (TiZrHf)1−x(AB)x and TiZrHf present nonuniform behaviors, and the ν of TiZrHf is lower than that of (TiZrHf)1−x(AB)x. Further, Figure 6e shows the θ D of (TiZrHf)1−x(AB)x as a function of concentration x, and the results show that the θ D linearly increases with the increasing concentration x, except for a few scattered points. Among these alloys, the θ D of (TiZrHf)1−x(NbMo)x (x = 0.33–0.89) and (TiZrHf)1−x(MoTa)x (x = 0.44–0.89) is larger than that of TiZrHf of 260.01 K. In Figure 6f, we show the Tm as a function of θ D for (TiZrHf)1−x(AB)x RHEAs, and it is seen that the Tm increases significantly with the increase of θ D , indicating that there is a strong positive correlation between Tm and θ D .
To further show the underlying effect of NbMo, NbTa, MoTa to TiZrHf on mechanical stability, the orbital Hamiltonian populations (COHP) [57,58] of (TiZrHf)0.44(AB)0.56, including pure TiZrHf, are calculated using the Local Orbital Basis Suite Towards Electronic-Structure Reconstruction (LOBSTER) program [59], and the COHP curves are plotted in Figure 7a–c. The red, blue, navy, and violet curves are, respectively, the COHP curves of TiZrHf, (TiZrHf)0.44(NbMo)0.5, (TiZrHf)0.44(NbTa)0.5, and (TiZrHf)0.44(MoTa)0.5. We focus on the chemical bonds of the first nearest neighbor with lengths less than 3.5 Å because the strengths of longer bonds are very weak. It can be found that, in pure TiZrHf, the chemical bonds of Ti-Zr, Ti-Hf, and Zr-Hf exhibit obvious antibonding states under the fermi level, indicating that the chemical bonds between them are very weak, and pure TiZrHf exhibits poor mechanical stability. When NbMo, NbTa, and MoTa are added to TiZHf, the chemical bonds of Nb-Mo, Nb-Ta, and Mo-Ta exhibit a strong bonding state, indicating that the addition of NbMo, NbTa, and MoTa significantly improves the mechanical stability.

4. Conclusions

The stability and mechanical properties of (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) refractory high-entropy alloys have been investigated by combining the first-principles with special quasi-random structure (SQS) method. Here are some main results as a function of concentration x, as follows:
(1)
The ΔHmix of (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) linearly decreases, whereas both ΔHmix and ΔSmix of (TiZrHf)1−x(NbTa)x increase initially and subsequently decrease, with the crossover occurring at x = 0.56. The ΔHmix of (TiZrHf)1−x(NbTa)x and (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) alloys are larger and lower than that of TiZrHf, respectively, while the ΔSmix of all (TiZrHf)1−x(AB)x is larger than that of TiZrHf.
(2)
The Ω of all (TiZrHf)1−x(AB)x (AB = NbMo, MoTa) first decreases sharply, followed by a gradual decrease. And the δ remains relatively stable around x = 0.56 for all cases, after which it decreases sharply until x = 0.89. The δ value of (TiZrHf)1−x(AB)x is higher than that of TiZrHf for x < 0.56 but becomes lower beyond this composition.
(3)
The VEC increases linearly from x = 0.11 to 0.89 while the ΔHf shows the opposite, except for (TiZrHf)0.67(NbTa)0.33. The VEC of all (TiZrHf)1−x(AB)x alloys increases, whereas their ΔHf decreases compared to that of TiZrHf.
(4)
The σp of (TiZrHf)1−x(AB)x increases linearly, reaching approximately 2.12 GPa, and the σp of all (TiZrHf)1−x(AB)x is larger than that of TiZrHf. B, E, and G also exhibit linear increases, and their values in all (TiZrHf)1−x(AB)x alloys are higher than those of TiZrHf, with some exceptions.
(5)
The C12C44 and G/B of all (TiZrHf)1−x(AB)x alloys increase, whereas the ν exhibits the opposite trend. Moreover, The C12C44 and G/B ratio of TiZrHf are lower and higher, respectively, than those of (TiZrHf)1−x(AB)x, and the ν of TiZrHf is lower than that of (TiZrHf)1−x(AB)x.

Author Contributions

Conceptualization, H.L., Y.Z. and T.F.; methodology, H.L., Z.R. and T.F.; software, Z.R. and T.F.; validation, Y.Z., T.F., T.H. and H.Y.; formal analysis, T.F.; investigation, H.L., Y.Z., Z.R. and T.F.; resources, Y.Z. and T.F.; data curation, Z.R. and T.H.; writing—original draft preparation, Z.R. and T.F.; writing—review and editing, T.F.; visualization, H.L., T.F., T.H. and H.Y.; supervision, T.F.; project administration, H.L. and T.F.; funding acquisition, H.L. and T.F. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (No. 52371009), the Xinjiang Development Program for Vanadium-Titanium Magnetite Concentration and Extraction Process Technology (No. 2022LQ01006), and the Research Project on Titanium Ore Smelting Process in Xinjiang (No. hmkjxm202206).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

Authors Yuanyuan Zhang and Touwen Fan were employed by the company Xinjiang Xianghe New Materials Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Miracle, D.B.; Senkov, O.N. A critical review of high entropy alloys and related concepts. Acta Mater. 2017, 122, 448–511. [Google Scholar] [CrossRef]
  2. Wang, S.; Wu, D.; She, H.; Wu, M.; Shu, D.; Dong, A.; Lai, H.; Sun, B. Design of high-ductile medium entropy alloys for dental implants. Mater. Sci. Eng. C 2020, 113, 110959. [Google Scholar] [CrossRef] [PubMed]
  3. Pink, E.; Eck, R. Refractory Metals and Their Alloys. In Materials Science and Technology: A Comprehensive Treatment, The Classic ed.; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar] [CrossRef]
  4. Praveen, S.; Kim, H.S. High-Entropy Alloys: Potential Candidates for High-Temperature Applications—An Overview. Adv. Eng. Mater. 2018, 20, 1700645. [Google Scholar] [CrossRef]
  5. Zhou, J.-L.; Cheng, Y.-H.; Chen, H.; Ma, K.; Wan, Y.-X.; Yang, J.-Y. Strengthening mechanisms and high-temperature oxidation properties of laser-clad TaNbZrTi refractory high entropy alloy coatings. J. Mater. Sci. 2023, 58, 16822–16840. [Google Scholar] [CrossRef]
  6. Zhao, Y.; Shi, W.; Huang, J. Microstructure evolution and performance effect of composite carbide (W,Ti)C enhanced Ni-based coatings fabricated by laser cladding. J. Mater. Res. Technol. 2025, 35, 298–307. [Google Scholar] [CrossRef]
  7. Yamabe-Mitarai, Y.; Yanao, K.; Toda, Y.; Ohnuma, I.; Matsunaga, T. Phase stability of Ti-containing high-entropy alloys with a bcc or hcp structure. J. Alloys Compd. 2022, 911, 164849. [Google Scholar] [CrossRef]
  8. Yeh, J.-W. Alloy Design Strategies and Future Trends in High-Entropy Alloys. JOM 2013, 65, 1759–1771. [Google Scholar] [CrossRef]
  9. Rogal, Ł.; Czerwinski, F.; Jochym, P.T.; Litynska-Dobrzynska, L. Microstructure and mechanical properties of the novel Hf25Sc25Ti25Zr25 equiatomic alloy with hexagonal solid solutions. Mater. Des. 2016, 92, 8–17. [Google Scholar] [CrossRef]
  10. Senkov, O.N.; Wilks, G.B.; Scott, J.M.; Miracle, D.B. Mechanical properties of Nb25Mo25Ta25W25 and V20Nb20Mo20Ta20W20 refractory high entropy alloys. Intermetallics 2011, 19, 698–706. [Google Scholar] [CrossRef]
  11. Qi, L.; Chrzan, D.C. Tuning Ideal Tensile Strengths and Intrinsic Ductility of bcc Refractory Alloys. Phys. Rev. Lett. 2014, 112, 115503. [Google Scholar] [CrossRef]
  12. Sheikh, S.; Shafeie, S.; Hu, Q.; Ahlström, J.; Persson, C.; Veselý, J.; Zýka, J.; Klement, U.; Guo, S. Alloy design for intrinsically ductile refractory high-entropy alloys. J. Appl. Phys. 2016, 120, 164902. [Google Scholar] [CrossRef]
  13. Dirras, G.; Lilensten, L.; Djemia, P.; Laurent-Brocq, M.; Tingaud, D.; Couzinié, J.P.; Perrière, L.; Chauveau, T.; Guillot, I. Elastic and plastic properties of as-cast equimolar TiHfZrTaNb high-entropy alloy. Mater. Sci. Eng. A 2016, 654, 30–38. [Google Scholar] [CrossRef]
  14. Rogal, L.; Ikeda, Y.; Lai, M.; Körmann, F.; Kalinowska, A.; Grabowski, B. Design of a dual-phase hcp-bcc high entropy alloy strengthened by ω nanoprecipitates in the Sc-Ti-Zr-Hf-Re system. Mater. Des. 2020, 192, 108716. [Google Scholar] [CrossRef]
  15. Zhang, L.; Fu, H.; Ge, S.; Zhu, Z.; Li, H.; Zhang, H.; Wang, A.; Zhang, H. Phase transformations in body-centered cubic NbxHfZrTi high-entropy alloys. Mater. Charact. 2018, 142, 443–448. [Google Scholar] [CrossRef]
  16. Ikeda, Y.; Gubaev, K.; Neugebauer, J.; Grabowski, B.; Körmann, F. Chemically induced local lattice distortions versus structural phase transformations in compositionally complex alloys. npj Comput. Mater. 2021, 7, 34. [Google Scholar] [CrossRef]
  17. Clatterbuck, D.M.; Krenn, C.R.; Cohen, M.L.; Morris, J.W. Phonon Instabilities and the Ideal Strength of Aluminum. Phys. Rev. Lett. 2003, 91, 135501. [Google Scholar] [CrossRef]
  18. Iijima, Y.; Nagase, T.; Matsugaki, A.; Wang, P.; Ameyama, K.; Nakano, T. Design and development of Ti–Zr–Hf–Nb–Ta–Mo high-entropy alloys for metallic biomaterials. Mater. Des. 2021, 202, 109548. [Google Scholar] [CrossRef]
  19. Yurchenko, N.; Panina, E.; Shaysultanov, D.; Zherebtsov, S.; Stepanov, N. Refractory high entropy alloy with ductile intermetallic B2 matrix/hard bcc particles and exceptional strain hardening capacity. Materialia 2021, 20, 101225. [Google Scholar] [CrossRef]
  20. Nagase, T.; Todai, M.; Wang, P.; Sun, S.-H.; Nakano, T. Design and development of (Ti, Zr, Hf)-Al based medium entropy alloys and high entropy alloys. Mater. Chem. Phys. 2022, 276, 125409. [Google Scholar] [CrossRef]
  21. Zhang, J.; Liu, K.; Fang, D.; Qiu, X.; Tang, D.; Meng, J. Microstructure, tensile properties, and creep behavior of high-pressure die-cast Mg–4Al–4RE–0.4 Mn (RE = La, Ce) alloys. J. Mater. Sci. 2009, 44, 2046–2054. [Google Scholar] [CrossRef]
  22. Mahmudi, R.; Kabirian, F.; Nematollahi, Z. Microstructural stability and high-temperature mechanical properties of AZ91 and AZ91 + 2RE magnesium alloys. Mater. Des. 2011, 32, 2583–2589. [Google Scholar] [CrossRef]
  23. Jin, L.; Kevorkov, D.; Medraj, M.; Chartrand, P. Al–Mg–RE (RE = La, Sm) systems: Thermodynamic evaluations and optimizations coupled with key experiments and Miedema’s model estimations. J. Chem. Thermodyn. 2013, 58, 166–195. [Google Scholar] [CrossRef]
  24. Sun, S.P.; Li, X.P.; Wang, H.J.; Jiang, H.F.; Yi, D.Q. First-principles investigations on the electronic properties and stabilities of low-index surfaces of L12-Al3Sc intermetallic. Appl. Surf. Sci. 2014, 288, 609–618. [Google Scholar] [CrossRef]
  25. Finnis, M.W. The theory of metal-ceramic interfaces. J. Phys. Condens. Matter 1996, 8, 5811–5836. [Google Scholar] [CrossRef]
  26. Wang, C.; Wang, C.Y. Ni/Ni3Al interface: A density functional theory study. Appl. Surf. Sci. 2009, 255, 3669–3675. [Google Scholar] [CrossRef]
  27. Zhou, W.F.; Ren, X.D.; Ren, Y.P.; Yuan, S.Q.; Ren, N.F.; Yang, X.Q.; Adu-Gyamfi, S. Initial dislocation density effect on strain hardening in FCC aluminium alloy under laser shock peening. Philos. Mag. 2017, 97, 917–929. [Google Scholar] [CrossRef]
  28. Zhao, S.J.; Stocks, G.M.; Zhang, Y.W. Stacking fault energies of face-centered cubic concentrated solid solution alloys. Acta Mater. 2017, 134, 334–345. [Google Scholar] [CrossRef]
  29. Dong, T.H.; Zhang, X.D.; Yang, L.M.; Wang, F. Effect of structural vacancies on lattice vibration, mechanical, electronic, and thermodynamic properties of Cr5BSi3. Chin. Phys. B 2022, 31, 026101. [Google Scholar] [CrossRef]
  30. Li, X. Phase stability and micromechanical properties of TiZrHf-based refractory high-entropy alloys: A first-principles study. Phys. Rev. Mater. 2023, 7, 113604. [Google Scholar] [CrossRef]
  31. Chen, S.-M.; Ma, Z.-J.; Qiu, S.; Zhang, L.-J.; Zhang, S.-Z.; Yang, R.; Hu, Q.-M. Phase decomposition and strengthening in HfNbTaTiZr high entropy alloy from first-principles calculations. Acta Mater. 2022, 225, 117582. [Google Scholar] [CrossRef]
  32. Meng, H.; Duan, J.-M.; Chen, X.-T.; Jiang, S.; Shao, L.; Tang, B.-Y. Influence of Local Lattice Distortion on Elastic Properties of Hexagonal Close-Packed TiZrHf and TiZrHfSc Refractory Alloys. Phys. Status Solidi 2021, 258, 2100025. [Google Scholar] [CrossRef]
  33. Hafner, J. Ab-initio simulations of materials using VASP: Density-functional theory and beyond. J. Comput. Chem. 2008, 29, 2044–2078. [Google Scholar] [CrossRef] [PubMed]
  34. Blöchl, P.E. Projected augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef] [PubMed]
  35. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533. [Google Scholar] [CrossRef]
  36. Feynman, R.P. Forces in Molecules. Phys. Rev. 1939, 56, 340. [Google Scholar] [CrossRef]
  37. Van de Walle, A.; Tiwary, P.; De Jong, M.; Olmsted, D.; Asta, M.; Dick, A.; Shin, D.; Wang, Y.; Chen, L.-Q.; Liu, Z.-K. Efficient stochastic generation of special quasirandom structures. Calphad 2013, 42, 13–18. [Google Scholar] [CrossRef]
  38. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  39. Nutor, R.K.; Cao, Q.; Wang, X.; Zhang, D.; Fang, Y.; Zhang, Y.; Jiang, J.-Z.; Selection, P. Phase Selection, Lattice Distortions, and Mechanical Properties in High-Entropy Alloys. Adv. Eng. Mater. 2020, 22, 2000466. [Google Scholar] [CrossRef]
  40. Takeuchi, A.; Inoue, A. Classification of Bulk Metallic Glasses by Atomic Size Difference, Heat of Mixing and Period of Constituent Elements and Its Application to Characterization of the Main Alloying Element. Mater. Trans. 2005, 46, 2817–2829. [Google Scholar] [CrossRef]
  41. Yang, X.; Zhang, Y. Prediction of high-entropy stabilized solid-solution in multi-component alloys. Mater. Chem. Phys. 2012, 132, 233–238. [Google Scholar] [CrossRef]
  42. Cheng, C.; Zhang, X.; Haché, M.J.R.; Zou, Y. Magnetron co-sputtering synthesis and nanoindentation studies of nanocrystalline (TiZrHf)x(NbTa)1−x high-entropy alloy thin films. Nano Res. 2022, 15, 4873–4879. [Google Scholar] [CrossRef]
  43. Guo, N.N.; Wang, L.; Luo, L.S.; Li, X.Z.; Chen, R.R.; Su, Y.Q.; Guo, J.J.; Fu, H.Z. Effect of composing element on microstructure and mechanical properties in Mo–Nb–Hf–Zr–Ti multi-principle component alloys. Intermetallics 2016, 69, 13–20. [Google Scholar] [CrossRef]
  44. Hobhaydar, A.; Wang, X.; Wang, Y.; Li, H.; Van Tran, N.; Zhu, H. Effect of tungsten doping on the irradiation resistance of FeCrV-based refractory medium entropy alloy for potential nuclear applications. J. Alloys Compd. 2023, 966, 171635. [Google Scholar] [CrossRef]
  45. Chen, L.; Li, Y.; Xiao, B.; Gao, Y.; Zhao, S. Chemical bonding, thermodynamic stability and mechanical strength of Ni3Ti/α-Al2O3 interfaces by first-principles study. Scr. Mater. 2021, 190, 57–62. [Google Scholar] [CrossRef]
  46. Han, F.; Yuan, M.; Wei, Z.; Yao, Y.; Yao, L.; Xin, L.; Shen, X. First-principles study of the Ti/Al3Ti interfacial properties. Appl. Surf. Sci. 2021, 544, 148960. [Google Scholar] [CrossRef]
  47. Klimashin, F.F.; Lobmaier, L.; Koutná, N.; Holec, D.; Mayrhofer, P.H. The MoN–TaN system: Role of vacancies in phase stability and mechanical properties. Mater. Des. 2021, 202, 109568. [Google Scholar] [CrossRef]
  48. Zhang, X.; Huang, Y. Mechanical, thermodynamic, and electronic studies on the Al3V/Al interface based on the density functional theory. Surf. Interfaces 2021, 25, 101212. [Google Scholar] [CrossRef]
  49. Otero-De-La-Roza, A.; Abbasi-Pérez, D.; Luaña, V. Gibbs2: A new version of the quasiharmonic model code. II. Models for solid-state thermodynamics, features and implementation. Comput. Phys. Commun. 2011, 182, 2232–2248. [Google Scholar] [CrossRef]
  50. Liu, T.; Ma, T.; Li, Y.; Ren, Y.; Liu, W. Stabilities, mechanical and thermodynamic properties of Al–RE intermetallics: A first-principles study. J. Rare Earths 2022, 40, 345–352. [Google Scholar] [CrossRef]
  51. Vasilyev, D.; Ikhsanov, R.S.; Zheleznyi, M.; Kartsev, A. Calculations of elastic and thermal properties of the strengthening C14 Fe6Nb4Al2 Laves phase using the density functional theory. J. Mater. Sci. 2025, 60, 5427–5441. [Google Scholar] [CrossRef]
  52. Benmakhlouf, A.; Daoud, S.; Bouarissa, N.; Allaoui, O. Elastic constants and thermophysical properties of CuPd: First-principles study. Rev. Mex. Física 2025, 71, 020501. [Google Scholar] [CrossRef]
  53. El Galta, A.; Masrour, R. First-principles calculations to investigate structural, magneto-electronic, elastic, thermodynamic, and thermoelectric properties of Co2LuY (Y = Zr, Hf) alloys for potential industrial application. Solid State Commun. 2025, 404, 116028. [Google Scholar] [CrossRef]
  54. Johnson, R.A. Analytic nearest-neighbor model for fcc metals. Phys. Rev. B 1988, 37, 3924–3931. [Google Scholar] [CrossRef]
  55. Fu, H.; Zhao, Z.; Liu, W.; Peng, F.; Gao, T.; Cheng, X. Ab initio calculations of elastic constants and thermodynamic properties of γ-TiAl under high pressures. Intermetallics 2010, 18, 761–766. [Google Scholar] [CrossRef]
  56. Frantsevich, I.N.; Voronov, F.F.; Bokuta, S.A. Elastic Constants and Elastic Moduli of Metals and Insulators Handbook; Naukova Dumka: Kiev, Ukraine, 1983. [Google Scholar]
  57. Dronskowski, R.; Bloechl, P.E. Crystal orbital Hamilton populations (COHP): Energy-resolved visualization of chemical bonding in solids based on density-functional calculations. J. Phys. Chem. 1993, 97, 8617–8624. [Google Scholar] [CrossRef]
  58. Deringer, V.L.; Tchougréeff, A.L.; Dronskowski, R. Crystal Orbital Hamilton Population (COHP) Analysis As Projected from Plane-Wave Basis Sets. J. Phys. Chem. A 2011, 115, 5461–5466. [Google Scholar] [CrossRef]
  59. Nelson, R.; Ertural, C.; George, J.; Deringer, V.L.; Hautier, G.; Dronskowski, R. LOBSTER: Local orbital projections; atomic charges, and chemical-bonding analysis from projector-augmented-wave-based density-functional theory. J. Comput. Chem. 2020, 41, 1931–1940. [Google Scholar] [CrossRef]
Figure 1. (a,b) The total energy of (TiZrHf)0.44(AB)0.56 (AB = NbMo, NbTa) as a function of cut-off energy.
Figure 1. (a,b) The total energy of (TiZrHf)0.44(AB)0.56 (AB = NbMo, NbTa) as a function of cut-off energy.
Coatings 15 01092 g001
Figure 2. The models of (TiZrHf)1−x(AB)x RHEAs.
Figure 2. The models of (TiZrHf)1−x(AB)x RHEAs.
Coatings 15 01092 g002
Figure 3. (a) The enthalpy of mixing ΔHmix, (b) entropy of mixing ΔSmix, (c) Ω, (d) standard deviation δ-parameter, (e) average valence electron concentration VEC and (f) formation enthalpy ΔHf of (TiZrHf)1−x(AB)x RHEAs as a function of x concentration. (g) The Evac of (TiZrHf)0.44(AB)0.56 containing pure TiZrHf.
Figure 3. (a) The enthalpy of mixing ΔHmix, (b) entropy of mixing ΔSmix, (c) Ω, (d) standard deviation δ-parameter, (e) average valence electron concentration VEC and (f) formation enthalpy ΔHf of (TiZrHf)1−x(AB)x RHEAs as a function of x concentration. (g) The Evac of (TiZrHf)0.44(AB)0.56 containing pure TiZrHf.
Coatings 15 01092 g003
Figure 4. (ac) The tensile stresses σ of (TiZrHf)1−x(AB)x RHEAs as a function of strains. (d) The peak stresses σp of (TiZrHf)1−x(AB)x RHEAs as a function of concentration x.
Figure 4. (ac) The tensile stresses σ of (TiZrHf)1−x(AB)x RHEAs as a function of strains. (d) The peak stresses σp of (TiZrHf)1−x(AB)x RHEAs as a function of concentration x.
Coatings 15 01092 g004
Figure 5. (a) The heat capacity CV, (b) entropy S, (c) thermal expansion coefficient αT and (d) Gibbs free energy Eg as a function of strains.
Figure 5. (a) The heat capacity CV, (b) entropy S, (c) thermal expansion coefficient αT and (d) Gibbs free energy Eg as a function of strains.
Coatings 15 01092 g005
Figure 6. (a) bulk modulus B, shear modulus G, Young’s modulus E, (b) Cauchy pressure C12C44, (c) Pugh’s G/B, (d) Poisson’s ratio ν, and (e) Debye temperature θ D of (TiZrHf)1−x(AB)x RHEAs as a function of x concentration. (f) The melting temperature Tm as a function of θ D for (TiZrHf)1−x(AB)x RHEAs.
Figure 6. (a) bulk modulus B, shear modulus G, Young’s modulus E, (b) Cauchy pressure C12C44, (c) Pugh’s G/B, (d) Poisson’s ratio ν, and (e) Debye temperature θ D of (TiZrHf)1−x(AB)x RHEAs as a function of x concentration. (f) The melting temperature Tm as a function of θ D for (TiZrHf)1−x(AB)x RHEAs.
Coatings 15 01092 g006
Figure 7. (ac) The COHP curves of (TiZrHf)0.44(AB)0.56.
Figure 7. (ac) The COHP curves of (TiZrHf)0.44(AB)0.56.
Coatings 15 01092 g007
Table 1. The calculated enthalpy of mixing ΔHmix (kJ/mol), entropy of mixing ΔSmix (J/(mol·K)), melting temperature Tm (K), Ω, δ-parameter (%), average valence electron concentration (VEC), and formation enthalpy ΔHf (kJ/mol).
Table 1. The calculated enthalpy of mixing ΔHmix (kJ/mol), entropy of mixing ΔSmix (J/(mol·K)), melting temperature Tm (K), Ω, δ-parameter (%), average valence electron concentration (VEC), and formation enthalpy ΔHf (kJ/mol).
AlloysΔHmixΔSmixTmΩδVECΔHf
TiZrHf09.132160.4811.674.005.40
(TiZrHf)0.89(NbMo)0.11−0.3411.662233.6177.1813.264.175.05
(TiZrHf)0.78(NbMo)0.22−0.7612.792306.7438.9613.924.333.89
(TiZrHf)0.67(NbMo)0.33−1.2613.302379.8725.1414.754.501.96
(TiZrHf)0.56(NbMo)0.44−1.8413.352453.0017.7614.024.671.51
(TiZrHf)0.44(NbMo)0.56−2.5112.972526.1313.0513.714.830.77
(TiZrHf)0.33(NbMo)0.67−3.2612.182599.269.7110.425.00−0.87
(TiZrHf)0.22(NbMo)0.78−4.0910.922672.397.138.295.17−2.54
(TiZrHf)0.11(NbMo)0.89−5.009.042745.524.964.995.33−4.92
(TiZrHf)0.89(NbTa)0.111.1211.662254.3423.4813.104.116.97
(TiZrHf)0.78(NbTa)0.221.9612.792348.1915.3313.014.227.42
(TiZrHf)0.67(NbTa)0.332.5213.302442.0412.9015.864.337.71
(TiZrHf)0.56(NbTa)0.442.8013.352535.8912.1014.624.445.95
(TiZrHf)0.44(NbTa)0.562.8012.972629.7412.1915.434.565.28
(TiZrHf)0.33(NbTa)0.672.5212.182723.5913.1711.484.673.39
(TiZrHf)0.22(NbTa)0.781.9610.922817.4515.7010.224.782.15
(TiZrHf)0.11(NbTa)0.891.129.042911.3023.515.454.890.52
(TiZrHf)0.89(MoTa)0.11−0.5211.662262.9550.4912.344.175.07
(TiZrHf)0.78(MoTa)0.22−1.0512.792365.4128.7213.284.334.60
(TiZrHf)0.67(MoTa)0.33−1.5913.302467.8720.6113.814.504.47
(TiZrHf)0.56(MoTa)0.44−2.1413.352570.3416.0314.604.671.93
(TiZrHf)0.44(MoTa)0.56−2.7012.972672.8012.8612.264.831.30
(TiZrHf)0.33(MoTa)0.67−3.2612.182775.2610.3711.345.00−0.70
(TiZrHf)0.22(MoTa)0.78−3.8310.922877.728.209.155.17−4.43
(TiZrHf)0.11(MoTa)0.89−4.419.042980.196.117.145.33−5.98
Table 2. The calculated elastic constants Cij (GPa) and Cauchy pressure C12C44 (GPa).
Table 2. The calculated elastic constants Cij (GPa) and Cauchy pressure C12C44 (GPa).
AlloysC11C12C44C12C44
TiZrHf155.6084.2040.8643.30
(TiZrHf)0.89(NbMo)0.11136.8691.6538.0453.61
(TiZrHf)0.78(NbMo)0.22152.0194.0637.9956.07
(TiZrHf)0.67(NbMo)0.33172.86100.0439.6060.44
(TiZrHf)0.56(NbMo)0.44190.43107.4039.2568.16
(TiZrHf)0.44(NbMo)0.56218.76110.8138.6372.19
(TiZrHf)0.33(NbMo)0.67251.02121.2742.4778.80
(TiZrHf)0.22(NbMo)0.78275.92136.4051.1085.30
(TiZrHf)0.11(NbMo)0.89321.77143.8259.6684.16
(TiZrHf)0.89(NbTa)0.11124.4896.1630.6665.49
(TiZrHf)0.78(NbTa)0.22138.7197.7333.2464.49
(TiZrHf)0.67(NbTa)0.33148.50102.3131.0671.25
(TiZrHf)0.56(NbTa)0.44165.37110.5138.3672.16
(TiZrHf)0.44(NbTa)0.56183.87113.1138.2374.88
(TiZrHf)0.33(NbTa)0.67204.55120.7341.4979.23
(TiZrHf)0.22(NbTa)0.78223.90130.2543.9686.28
(TiZrHf)0.11(NbTa)0.89247.21136.9444.1892.76
(TiZrHf)0.89(MoTa)0.11126.7097.2238.1259.10
(TiZrHf)0.78(MoTa)0.22152.3897.0741.7755.30
(TiZrHf)0.67(MoTa)0.33177.5698.8442.5756.28
(TiZrHf)0.56(MoTa)0.44202.47108.4248.0360.39
(TiZrHf)0.44(MoTa)0.56218.00118.6746.1572.52
(TiZrHf)0.33(MoTa)0.67254.04130.6251.6578.97
(TiZrHf)0.22(MoTa)0.78289.34142.8059.4183.39
(TiZrHf)0.11(MoTa)0.89324.49155.1065.7089.40
Table 3. The calculated bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Pugh’s G/B, and Poisson’s ratio ν.
Table 3. The calculated bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Pugh’s G/B, and Poisson’s ratio ν.
AlloysBGEG/BυϴD
TiZrHf108.0038.71103.740.3580.340260.01
(TiZrHf)0.89(NbMo)0.11106.7230.8784.470.2890.368233.74
(TiZrHf)0.78(NbMo)0.22113.3834.0992.950.3010.363245.99
(TiZrHf)0.67(NbMo)0.33124.3138.29104.180.3080.360261.00
(TiZrHf)0.56(NbMo)0.44135.0840.14109.560.2970.365267.96
(TiZrHf)0.44(NbMo)0.56146.7944.18120.440.3010.363281.68
(TiZrHf)0.33(NbMo)0.67164.5250.36137.080.3060.361301.40
(TiZrHf)0.22(NbMo)0.78182.9157.90157.110.3170.357323.88
(TiZrHf)0.11(NbMo)0.89203.1470.05188.490.3450.345356.87
(TiZrHf)0.89(NbTa)0.11105.6022.4962.990.2130.401196.34
(TiZrHf)0.78(NbTa)0.22111.3927.3875.910.2460.386212.33
(TiZrHf)0.67(NbTa)0.33117.7127.5876.750.2340.391209.42
(TiZrHf)0.56(NbTa)0.44128.8033.5392.570.2600.380226.71
(TiZrHf)0.44(NbTa)0.56136.7037.07101.980.2710.376234.25
(TiZrHf)0.33(NbTa)0.67148.6741.66114.310.2800.372244.26
(TiZrHf)0.22(NbTa)0.78161.4645.09123.740.2790.372250.10
(TiZrHf)0.11(NbTa)0.89173.7048.28132.550.2780.373254.90
(TiZrHf)0.89(MoTa)0.11107.0526.0572.270.2430.387210.56
(TiZrHf)0.78(MoTa)0.22115.5035.4096.370.3070.361239.44
(TiZrHf)0.67(MoTa)0.33125.0841.25111.500.3300.351252.94
(TiZrHf)0.56(MoTa)0.44139.7747.63128.300.3410.347266.39
(TiZrHf)0.44(MoTa)0.56151.7847.52129.100.3130.358261.59
(TiZrHf)0.33(MoTa)0.67171.7655.47150.230.3230.354277.22
(TiZrHf)0.22(MoTa)0.78191.6564.61174.260.3370.348293.83
(TiZrHf)0.11(MoTa)0.89211.5672.74195.770.3440.346306.37
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Luo, H.; Zhang, Y.; Ruan, Z.; Fan, T.; Hu, T.; Yan, H. Theoretical Approach of Stability and Mechanical Properties in (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) Refractory High-Entropy Alloys. Coatings 2025, 15, 1092. https://doi.org/10.3390/coatings15091092

AMA Style

Luo H, Zhang Y, Ruan Z, Fan T, Hu T, Yan H. Theoretical Approach of Stability and Mechanical Properties in (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) Refractory High-Entropy Alloys. Coatings. 2025; 15(9):1092. https://doi.org/10.3390/coatings15091092

Chicago/Turabian Style

Luo, Heng, Yuanyuan Zhang, Zixiong Ruan, Touwen Fan, Te Hu, and Hongge Yan. 2025. "Theoretical Approach of Stability and Mechanical Properties in (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) Refractory High-Entropy Alloys" Coatings 15, no. 9: 1092. https://doi.org/10.3390/coatings15091092

APA Style

Luo, H., Zhang, Y., Ruan, Z., Fan, T., Hu, T., & Yan, H. (2025). Theoretical Approach of Stability and Mechanical Properties in (TiZrHf)1−x(AB)x (AB = NbTa, NbMo, MoTa) Refractory High-Entropy Alloys. Coatings, 15(9), 1092. https://doi.org/10.3390/coatings15091092

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop