Next Article in Journal
Research Progress and Challenges in Laser-Controlled Cleaning of Aluminum Alloy Surfaces
Next Article in Special Issue
The Effect of Fe Addition in the RM(Nb)IC Alloy Nb–30Ti–10Si–2Al–5Cr–3Fe–5Sn–2Hf (at.%) on Its Microstructure, Complex Concentrated and High Entropy Phases, Pest Oxidation, Strength and Contamination with Oxygen, and a Comparison with Other RM(Nb)ICs, Refractory Complex Concentrated Alloys (RCCAs) and Refractory High Entropy Alloys (RHEAs)
Previous Article in Journal
Influence of Scale Effect on Strength and Deformation Characteristics of Rockfill Materials
Previous Article in Special Issue
Characterisation of a Novel Complex Concentrated Alloy for Marine Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vacancy Energetics and Diffusivities in the Equiatomic Multielement Nb-Mo-Ta-W Alloy

1
Department of Materials Science and Engineering, University of California, Los Angeles, CA 90095, USA
2
Department of Mechanical and Aerospace Engineering, University of California, Los Angeles, CA 90095, USA
*
Author to whom correspondence should be addressed.
Materials 2022, 15(15), 5468; https://doi.org/10.3390/ma15155468
Submission received: 22 June 2022 / Revised: 2 August 2022 / Accepted: 3 August 2022 / Published: 8 August 2022
(This article belongs to the Special Issue Compositional Complex Alloys: From Amorphous to High-Entropy)

Abstract

:
In this work, we study vacancy energetics in the equiatomic Nb-Mo-Ta-W alloy, especially vacancy formation and migration energies, using molecular statics calculations based on a spectral neighbor analysis potential specifically developed for Nb-Mo-Ta-W. We consider vacancy properties in bulk environments as well as near edge dislocation cores, including the effect of short-range order (SRO) by preparing supercells through Metropolis Monte-Carlo relaxations and temperature on the calculation. The nudged elastic band (NEB) method is applied to study vacancy migration energies. Our results show that both vacancy formation energies and vacancy migration energies are statistically distributed with a wide spread, on the order of 1.0 eV in some cases, and display a noticeable dependence on SRO. We find that, in some cases, vacancies can form with very low energies at edge dislocation cores, from which we hypothesize the formation of stable ‘superjogs’ on edge dislocation lines. Moreover, the large spread in vacancy formation energies results in an asymmetric thermal sampling of the formation energy distribution towards lower values. This gives rise to effective vacancy formation energies that are noticeably lower than the distribution averages. We study the effect that this phenomenon has on the vacancy diffusivity in the alloy and discuss the implications of our findings on the structural features of Nb-Mo-Ta-W.

1. Introduction

High-entropy alloys (HEAs), also known as multicomponent alloys or multiprincipal element alloys, consist of four or more distinct elements in equimolar or near-equimolar proportions [1,2,3,4,5]. Some HEAs are observed to exhibit combinations of desirable properties which are not common in conventional alloys, such as good irradiation resistance, corrosion resistance, thermal stability, and high strength [6,7,8], especially under harsh working environments. This makes them attractive candidate materials for applications in the energy sector, the aerospace industry, and transportation industries [9,10,11,12]. Their extraordinary properties and great potential for a variety of applications have stimulated research to investigate the underlying mechanisms [13,14,15,16,17,18]. Among the several different classes of HEA [7], refractory multi-element alloys (RMEAs) are composed of up to four or more refractory transition metals from group IVB, VB, VIB [14,19]. As their name indicates, RMEAs are attractive for their outstanding mechanical properties at elevated temperatures. They are able to retain a single phase bcc structure and have well-balanced combinations of high strength-weight ratio, fracture toughness, and ductility up to very high temperatures [20,21,22,23,24,25]. It is thought that the limited self-diffusion reported for most RMEAs plays a key role in the outstanding thermal stability and mechanical properties at high temperatures.
At high temperatures, vacancy properties are known to control RMEA performance such as the irradiation resistance, phase stability and creep deformation [26,27,28,29]. Thus, gaining a detailed understanding of vacancy energetics is a first important step to quantify the parameters that govern mass transport properties in these alloys. Of particular interest is the quantification of the (self) diffusion coefficient, which is universally reported to be very low for RMEA such as Nb-Mo-Ta-W [30,31,32]. Moreover, it is important to study the implications of high configurational entropy on defect properties such as those of vacancies, both structurally and also in terms of the statistical significance of having a large spread in the energy distributions.
For example, many works have shown that short-range order (SRO) can have significant effects on the mechanical, electronic, magnetic, and irradiation properties of HEAs [33,34,35,36,37,38], as well as vacancy energetics [39,40,41]. Thus, to gain a more comprehensive understanding of the interactions between chemical lattice complexity and vacancy energetics in HEAs, an investigation into the effect of SRO on vacancy energetics is of the essence.
Further, with vacancy formation energies in RMEAs presumed to be as high as in the pure elemental refractory elements, it is expected that heterogeneous nucleation at intrinsic defects, such as dislocations and grain boundaries, will be the dominant source of vacancies [41,42]. Thus, here we also consider vacancy energetics in the proximity of edge dislocation cores.
While it is well known that refractory metals display high vacancy formation energies, in multi-principal element alloys bulk formation energies are often distributed across a wide energy range [43,44,45,46]. This introduces the possibility of energies significantly lower than the distribution averages [43,44], opening the door to enhanced vacancy-mediated behavior compared to pure refractory metals. Moreover, since vacancies are generally produced at material heterogeneities, it is of interest to investigate the differences between formation energy distributions in the bulk and at selected material defects such as dislocations, grain boundaries, and free surfaces. While bulk vacancy formation distributions have been calculated for selected refractory multi-principal element alloys [46,47,48], here we also investigate vacancy formation at edge dislocation cores and study the implications on diffusion and dislocation climb of broadly-distributed vacancy energies in Nb-Mo-Ta-W. Furthermore, we study vacancy migration in the bulk and at the edge dislocation core to understand the effect that chemical composition has on these properties. To our knowledge, this has not been done before.
The paper is organized as follows. First, we discuss the atomistic simulation method, including the details of construction of the random Nb-Mo-Ta-W model, the construction of supercells with SRO, and the methods of calculating vacancy formation and migration energies, E V f and E V m , in a variety of scenarios. Then, we present results for E V f and E V m as a function of temperature by adopting supercell constrictions with the appropriate SRO at each temperature. Moreover, we report on the results of the same calculations but in the presence of an edge dislocation. This is followed by a discussion to provide detailed analysis of the results and demonstrate how the results may be used to calculate the diffusion coefficient in the alloy. Finally, we end the paper with our main conclusions.

2. Methods

2.1. Molecular Statics Calculations

All molecular statics calculations were performed using the LAMMPS code [49] with an interatomic potential adopted in this work developed by Li et al. [50] for the Nb-Mo-Ta-W system. The relaxed configurations were visualized using OVITO [51]. Initially, cells with a random distribution of Nb, Mo, Ta, and W in equal proportions were generated. The cells were oriented along x = [ 111 ] , y = [ 1 1 ¯ 0 ] , z = [ 11 2 ¯ ] in a body-centered cubic (BCC) lattice with dimensions of 49.6 × 43.5 × 41 Å. Periodic boundary conditions were enforced in all three dimensions. The total number of atoms inside the simulation cell ranged from 5600 to 5760.

2.2. Metropolis Monte-Carlo Method

To study the effect of SRO on vacancy energetics, we prepare simulation cells with a random distribution of atomic species and apply the Metropolis Monte-Carlo method to generate optimized atomic configurations at different temperatures. The Monte-Carlo relaxations follow the sequence listed below:
1.
Randomly choose a type of atom-pair with different atom species (e.g., Nb and Ta, Nb and Mo), and attempt swaps for 100 times for candidates that belong to the chosen type.
2.
Run a conjugate gradient (CG) minimization of the structure at that point.
3.
Repeat steps 1 and 2 until the total energy of the supercell converges with the number of iterations.
A series of temperatures ranging from 300 K to 1500 K is chosen and we then quantify the degree of SRO as determined by a pre-specified order parameter (discussed below).

2.3. Vacancy Formation and Migration Energies

The vacancy formation energy, E V f , is calculated as follows:
E V f = E N 1 E N + μ avg
where E N 1 and E N are the total energies of relaxed supercells with and without a vacancy, respectively, and  μ avg is the chemical potential energy of the average alloy computed as:
μ avg = 1 4 ( μ Nb + μ Mo + μ Ta + μ W )
where μ i ( i = Nb , Mo , Ta , W ) is the chemical potential of a pure element calculated in the context of its pure metal.
Since E V f varies due to the change of local chemical environment in the alloy, up to 300 different lattice sites were randomly chosen for each cell to collect vacancy formation energies in as many local environments as possible. For the purpose of comparison, the vacancy formation energies in pure constituents of the alloy and the ‘average’ alloy (average across all four constituents) are also computed and analyzed.
For their part, vacancy migration energies, E V m , are calculated using the nudged elastic band (NEB) method implemented in LAMMPS. Vacancy trajectories along the first nearest neighbor direction [ 111 ] were subdivided into a total number of eight replicas, including six intermediate replicas plus the initial and final configurations. The NEB method furnishes the minimum energy path (MEP) from the initial to the final replica. After finding the MEP, the  E V m is computed as the difference between the energy at the saddle point and the energy of one of the end replicas. Similar to the case of E V f , we try over 150 different lattice locations to capture the configurational variations introduced by the alloy compositional fluctuations.

3. Results

3.1. Equilibrium Lattice Parameter

To calculate the lattice parameter a 0 , we ran conjugate gradient minimizations of supercells containing 16,000 atoms arranged into a random structure. Figure 1 shows the energy per atom as a function of the unit cell size, a. The error bars correspond to 20 independently generated random configurations for each value of a. A second-degree polynomial fit to the data yields a value of a 0 = 3.243 Å.
Using the same method, we simulated the lattice parameters of the individual alloy constituents as well as their weighted average given in Table 1. As the results show, a 0 for Nb and Ta are 2.5% larger than the average value, while the values for Mo and W are 2.1% smaller.

3.2. Short Range Order Calculation

The degree of SRO is quantified by the Warren–Cowley parameter [52], defined as:
η α β = 1 i N β x i α β N β c β
where x α β is the local concentration of atoms of type α neighboring an atom of type β (here we consider only the first-nearest neighbor (1NN) shells), while c β is the global concentration of atom species β in the cell. N β is the number of atoms of type β in the computational cell. With this definition, there are 10 unique pairs, characterized by the following values:
η α β < 0 , ordering = 0 , random > 0 . separation
The Warren–Cowley parameters of the system are computed as a function of temperature after Metropolis Monte-Carlo relaxations of computational cells consisting of 5760 atoms. Figure 2 shows the evolution of the supercell energy as a function of temperature. Convergence is achieved within 150,000 iterations at the highest temperatures, while it requires in excess of 600,000 steps below 900 K. The values for the 10 distinct pairs of elements are plotted in Figure 3 as a function of temperature. The small band around zero (‘error band’) marks the spontaneous SRO introduced by species fluctuations in the random configuration. In general, pairs of elements from the same group, either group VB or group VIB (e.g., Ta-Nb and Mo-W), tend to segregate from one another ( η α β > 0 ), while pairs composed of elements from different groups (e.g., Ta-Mo, and Nb-W) show a tendency toward ordering ( η α β < 0 ). Among these, the pair Ta-Mo becomes the most negative one, indicating the strongest attraction between the two elements involved. Our calculations agree with results of previous SRO studies in the equimolar Mo-Nb-Ta-V-W system [53,54]. Interestingly, at 300 K, the Warren–Cowley parameter of Nb-Mo decreases slightly while that of Ta-Mo continues to increase, which suggests that a competition may exist between Ta and Nb when pairing with Mo.

3.3. Vacancy Energetics in Bulk Systems

3.3.1. Vacancy Formation Energies in Bulk Nb-Mo-Ta-W

The chemical potential energies used in the calculation of vacancy formation energies in the real alloy are displayed in Table 2.
To calculate the vacancy formation energy distributions, we collect data for up to 300 randomly distributed vacancies in each box with the corresponding SRO as obtained in the previous section. The E V f distributions of formation energies are shown in Figure 4. The dashed line in each figure represents the statistical mean of each distribution, which is listed in Table 3.
Our results show that the effect of SRO on E V f is relatively weak, and that as the temperature increases, the random alloy becomes a good representative of the relaxed supercells with SRO. The vacancy formation energies of the constituent metals, as well as their arithmetic averages, are also presented in Table 3. Interestingly, the average of the individual constituent elements is noticeably larger than those of the actual alloy.

3.3.2. Effective Vacancy Formation Energy in Bulk

While the vacancy formation energy distributions given in Figure 4 represent the spectrum of possibilities for vacancies to form in a random lattice, in reality the thermal vacancy concentration is not well predicted by the distribution averages. This is because thermal sampling of the distributions disproportionately favors low-energy occurrences. As such, it is of interest to calculate the equilibrium concentration of vacancies as a function of temperature, C V ( T ) , by direct sampling of Figure 4.
Next, we visit a large number of lattice sites, N = 10 10 , and evaluate the Boltzmann probability of forming a vacancy by uniformly sampling the distributions given in Figure 4:
p i = exp E i f k T
where E i f is the i th sampling of the vacancy-formation energies, and k is the Boltzmann constant. C V ( T ) is obtained as the ratio of the number of successful vacancy formation events, n ( T ) , defined by:
ξ i < p i
where ξ i is a uniform random number between 0 and 1, and the total number of trials is N, i.e.,:
C V ( T ) = n ( T ) N
One then matches the calculated value of C V ( T ) to an ‘effective’ formation energy:
E eff f ( T ) = k T log n ( T ) N
The effective formation energies at temperatures ranging from 1500 K to 2000 K are shown in Figure 5. The values are listed in Table 4. The value of E eff f starts from about 2.37 eV at 1500 K, and grows with a decreasing rate to approximately 2.40 eV at 2000 K. It is clear in the figure that the sensitivity of E eff f to temperature decreases beyond 1500 K which leads to slow increase in E eff f , resulting in lower effective energies at all temperatures compared with the average E V f of the distribution in the random bulk. The significance of these calculations will be explained in the discussion session.

3.3.3. Vacancy Migration Energies in the Bulk

The MEP of vacancy migration was obtained by the NEB method for all the pure elements and for the equiatomic Nb-Mo-Ta-W system. A total of 150 different paths were analyzed for the alloy. The results are shown in Figure 6. As expected, the reaction path in the alloy is not symm etric, due to chemical energy differences between the initial and final states. As a result, two distinct barriers can be extracted from the alloy.
As in the previous section, the distributions of vacancy migration energy, E V m , of cells with and without SRO are shown in Figure 7. The mean values of E V m for all temperatures are given in Table 5. Here, a clear decrease in E V m can be seen with decreasing SRO.
To compare, the migration energies of pure components and the ’average’ alloy are simulated and displayed in Table 5 as well. Among all the components, W has the highest migration energy of 2.074 eV, which is higher than all Nb-Mo-Ta-W alloys, while other pure elements have energies close to (e.g., Mo), or lower than (e.g., Nb and Ta) the Nb-Mo-Ta-W alloys. The average of the four pure metals is 1.529 eV, smaller than the average migration energies of all the Nb-Mo-Ta-W alloys.
In the bcc crystal lattice, there are eight 1NN jumps originating from the same site. It is thus important to check if vacancy migration barriers are correlated to the originating site, or completely uncorrelated. To that end, we compute barriers purely randomly (i.e., select a lattice site at random and one 1NN path at random) or in a correlated fashion (one lattice site at random and all eight paths at that location). The results for both cases are shown in Figure 8, from which we see that there is no distinct difference between the energy distributions of the single direction and the eight-direction case. This suggests that the migration barriers are truly uncorrelated from the originating lattice site.

3.4. Vacancy Energetics at Edge Dislocation Cores

Formation Energies

Vacancy formation is significantly facilitated by heterogeneities. A good example of structural heterogeneity is an edge dislocation. To this end, we calculate the vacancy formation energies as a function of distance to an edge dislocation core on the atomic plane of the compressive region immediately adjacent to the glide plane. Figure 9 shows a relaxed 1 2 [ 111 ] edge dislocation dipole, with atoms colored by their chemical nature, created as explained in the paper by Hossain et al. [55]. The figure also shows the locations where vacancy formation energies have been calculated.
Figure 10 shows the results obtained from 300 independent evaluations. As shown, the vacancy formation energies are drastically reduced within four Burgers vector distances from the bulk values around 2.5 eV down to 1.0 eV at position ‘0’.
As before, SRO only has a marginal effect on these results. The specific values at location ‘0’ as a function of SRO at different temperatures and also for the elemental metals are given in Table 6. The formation energies for the Nb-Mo-Ta-W system are higher than those of the pure metals and the ’average’ alloy.

4. Discussion

4.1. Statistical Distribution of Vacancy Energetics

As displayed in Figure 4 and Figure 7, we find that both vacancy formation energies and migration energies are statistically distributed, which agrees with observations in recent works in HEAs [29,43,56,57]. The large spread in the distributions can be explained by the variety of local chemical environments in the alloy. Our results also demonstrate that the spread width and average energies of both distributions of E V f and E V m display a minor but non-negligible dependence on SRO.

4.2. Analysis of Vacancy Formation Energy and Migration Energy

Since the formation energy distributions are quite widely spread, vacancies with large formation energies are rarely activated, resulting in the asymmetric thermal sampling towards low formation energies. Therefore, the effective formation energies should be lower than the statistical averages of the formation energy distributions, which is confirmed by comparing the values of E V f (see Table 3) and E eff f (see Table 4). As temperature decreases, E eff f decreases with an increasing rate because the possibility of sampling vacancies with large barriers drops exponentially.
The presence of edge dislocation core effectively changes the formation energy barriers of vacancies within a few Burgers vector distances from the core, and its effect becomes negligible once vacancies are beyond five Burgers vector distances from the dislocation core (see Figure 10). The formation barriers are reduced to below 1.0 eV on average at core in the compressive region (see Figure 9), with occasional occurrence of near-zero E V f .
Unlike E V f , the statistical averages of distributions of E V m adequately describe the barriers of migration. During migration, a vacancy can visit a variety of chemical environments over a long distance, which involves a great number of independent samplings of E V m with varying values. Therefore, the value of stochastic average barrier of migration by sampling is close to the statistical average computed from the migration energy distribution.

4.3. Self-Diffusion Coefficients of Nb-Mo-Ta-W

Refractory high-entropy-alloys are generally said to display sluggish diffusion [58,59]. While this may be logically inferred from the known self-diffusion coefficients in the parent metals, here we have the opportunity to provide a direct assessment based on the results presented in this paper.
We start with the general Arrhenius expression for the diffusivity of a substitutional atomic species i:
D i = D 0 i exp Δ H i k T
where Δ H i is the activation enthalpy and D 0 i is a temperature-independent diffusion pre-factor. Δ H i is calculated as the sum of the formation and migration energies of a vacancy:
Δ H i = E f i + E m i
For Nb-Mo-Ta-W, we use the effective formation energies as the source for the values of E f i (from Figure 5) since they describe the true thermal concentrations of vacancies in the alloy. In terms of E m i , the statistical averages do adequately represent the migration barrier (see Table 5). In particular, as we demonstrated in Section 3.3.3, the vacancy migration energies can be considered uncorrelated with the originating jump site, which simplifies the sampling of the distributions given in Figure 8. The pre-exponential factors are computed using the expression:
D 0 = 1 6 z f ν 0 b 2
where z = 8 is the 1NN coordination number in bcc lattices, f 0.75 is a correlation factor, b = 3 2 a 0 is the 1NN jump distance, and a 0 for different systems can be found in Table 1. The frequency ν 0 is the Debye frequency, which is computed as:
ν 0 3 = 6 π 2 n c 3
where n is the atomic density, c is the velocity of sound in the material, expressed as:
n = 2 a 0 3
c = K ρ
where K is the shear modulus, and ρ is the mass density, computed using the equation:
ρ = 2 m N A a 0 3
where m is the atomic weight, 2 represents the number of atoms per unit cell in bcc structure, and N A = 6.022 × 10 23 is Avogadro’s constant. Values of D 0 and parameters used for computing the diffusivity of each system are listed in Table 7. At the moment, we do not have parameters in Equation (10) specific for the alloy, and thus we simply use a simple average of the values of the pure constituent metals. As such, the relevance of D 0 for this study is very marginal and is only used for estimation purposes.
Figure 11 shows an Arrhenius plot of the diffusivities of the elemental metals and the Nb-Mo-Ta-W system for both the effective vacancy concentrations and the distribution averages.
The figure clearly shows that the diffusivities of the pure metal and the alloy (in the two alternatives considered) are extremely low, thus substantiating the notion of sluggish diffusion typically attributed to RHEA.

4.4. Comparison with Other Works

In a recent molecular dynamics study by Luke [48], the average formation energy of a vacancy in random Nb-Mo-Ta-W was given as 2.26 eV, which is consistent with our result of 2.48 eV. Similar values are also obtained in ternary equiatomic V-Ta-W [61], which display an average value of 2.81 eV. In another work by Byggmastar et al. [54], an average formation energy of 3.3 eV was obtained in Mo-Nb-Ta-V-W using density-functional theory (DFT) calculations, and 3.1 eV using the Gaussian approximation potential framework. Using DFT, Roy et al. [62] studied vacancy stability in the body-centered cubic ( Mo 0.95 W 0.05 ) 0.85 Ta 0.10 ( TiZr ) 0.05 and computed the average formation energies ranging from 3.4 eV to 3.52 eV by alternating chemical environment. All their observations agree with our findings about the effect of SRO on vacancy energetics, i.e., the variation of E V f and E V m as a function of the local environment.
As for vacancy migration, a recent work of the ternary equiatomic bcc Mo-Nb-Ta alloy by Xing et al. [46] shows that, when SRO is introduced into the system, the chemical ordering induces a localizing trapping effect which increases the migration energies. This is in agreement with our results (Figure 7). They also report a broadening with SRO of the migration energy distributions, which in contrast is not so clear-cut in our case.
A study by Wang et al. [63] shows that the equilibrium concentrations of vacancies are greatly enhanced by the high entropy in the HEAs compared to pure metals, which corresponds to our finding that the effective formation energy of Nb-Mo-Ta-W is smaller than the average value of E V f obtained from its pure constituents especially at temperatures lower than 1600 K.

5. Conclusions

We end the paper with our most important conclusions.
1.
Both vacancy formation and migration energies in Nb-Ta-Mo-W are defined by statistical distributions with a wide spread, on the order of 1.0 eV in some cases.
2.
Vacancy energetics in Nb-Mo-Ta-W display a non-negligible dependence on SRO, which is reflected by the decrease in E V f from 2.54 eV to 2.48 eV and E V m from 1.89 eV to 1.71 eV as SRO weakens with increasing temperature.
3.
The vacancy formation energies are reduced by 1.4 eV on average as they approach an edge dislocation core from the bulk. Vacancies with low energies near zero can be found at core positions, from which we hypothesize that the formation of ‘superjogs’ on edge dislocation lines would be easy.
4.
Due to the widespread distribution of vacancy formation energies, its thermal sampling becomes asymmetric towards lower values, resulting in lowering in effective vacancy formation energies compared to statistical averages in formation energy distributions.
5.
The effective diffusivity in Nb-Mo-Ta-W using E eff f is smaller than the diffusivities of Nb and Ta, however it is larger than those of Nb-Mo-Ta-W using the statistical average E V f as well as Mo and W, the value of which starts from 10 19 m 2 · s 1 at 1500 K to 10 16 m 2 · s 1 at 2000 K, confirming the sluggish diffusion reported in the equimolar Nb-Ta-Mo-W alloy.

Author Contributions

Conceptualization, J.M.; Data curation, X.Z. and S.H.; Formal analysis, X.Z. and J.M.; Funding acquisition, J.M.; Investigation, X.Z. and S.H.; Methodology, X.Z. and J.M.; Project administration, J.M.; Resources, J.M.; Software, X.Z.; Supervision, J.M.; Validation, J.M.; Visualization, X.Z.; Writing—original draft, X.Z.; Writing—review & editing, X.Z. and J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been funded by the National Science Foundation under Grant No. DMR-1611342.

Data Availability Statement

The raw data generated in the work described in this article can be made available upon reasonable request to the authors.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
HEAHigh Entropy Alloy
RMEARefractory Multi-element Alloy
LAMMPSLarge-scale Atomic/Molecular Massively Parallel Simulator
NEBNudged Elastic Band
MEPMinimum Energy Path
CGConjugate Gradient
SROShort-range Order
1NN1st-nearest Neighbor

References

  1. Cantor, B.; Chang, I.; Knight, P.; Vincent, A. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375, 213–218. [Google Scholar] [CrossRef]
  2. Yeh, J.-W.; Chen, S.-K.; Lin, S.-J.; Gan, J.-Y.; Chin, T.-S.; Shun, T.-T.; Tsau, C.-H.; Chang, S.-Y. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  3. Kozak, R.; Sologubenko, A.; Steurer, W. Single-phase high-entropy alloys: An overview. Z. Krist.-Cryst. Mater. 2015, 230, 55–68. [Google Scholar] [CrossRef] [Green Version]
  4. Gao, M.C.; Jablonski, P.D.; Hawk, J.A.; Alman, D.E. High-entropy alloys: Formation and properties. In Proceedings of the ASME 2018 Symposium on Elevated Temperature Application of Materials for Fossil, Nuclear, and Petrochemical Industries, Seattle, DC, USA, 3–5 April 2018; p. V001T01A004. [Google Scholar]
  5. Senkov, O.; Miller, J.; Miracle, D.; Woodward, C. Accelerated exploration of multi-principal element alloys for structural applications. Calphad 2015, 50, 32–48. [Google Scholar] [CrossRef]
  6. Li, Z.; Pradeep, K.G.; Deng, Y.; Raabe, D.; Tasan, C.C. Metastable high-entropy dual-phase alloys overcome the strength-ductility trade-off. Nature 2016, 534, 227. [Google Scholar] [CrossRef]
  7. 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] [Green Version]
  8. Tsai, M.H.; Yeh, J.W. High-entropy alloys: A critical review. Mater. Res. Lett. 2014, 2, 107–123. [Google Scholar] [CrossRef]
  9. Xia, S.Q.; Zhen, W.; Yang, T.F.; Zhang, Y. Irradiation behavior in high entropy alloys. J. Iron Steel Res. 2015, 22, 879–884. [Google Scholar] [CrossRef]
  10. Xia, S.; Gao, M.C.; Yang, T.; Liaw, P.K.; Zhang, Y. Phase stability and microstructures of high entropy alloys ion irradiated to high doses. J. Nucl. Mater. 2016, 480, 100–108. [Google Scholar] [CrossRef] [Green Version]
  11. Kumar, N.K.; Li, C.; Leonard, K.; Bei, H.; Zinkle, S. Microstructural stability and mechanical behavior of fenimncr high entropy alloy under ion irradiation. Acta Mater. 2016, 113, 230–244. [Google Scholar] [CrossRef] [Green Version]
  12. Egami, T.; Guo, W.; Rack, P.; Nagase, T. Irradiation resistance of multicomponent alloys. Metall. Mater. Trans. 2014, 45, 180–183. [Google Scholar] [CrossRef]
  13. Senkov, O.N.; Senkova, S.V.; Woodward, C.J.A.M. Effect of aluminum on the microstructure and properties of two refractory high-entropy alloys. Acta Mater. 2014, 68, 214–228. [Google Scholar] [CrossRef]
  14. 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]
  15. Juan, C.C.; Tsai, M.H.; Tsai, C.W.; Lin, C.M.; Wang, W.R.; Yang, C.C.; Chen, S.K.; Lin, S.J.; Yeh, J.W. Enhanced mechanical properties of HfMoTaTiZr and HfMoNbTaTiZr refractory high-entropy alloys. Intermetallics 2015, 62, 76–83. [Google Scholar] [CrossRef]
  16. Senkov, O.N.; Woodward, C.; Miracle, D.B. Microstructure and properties of aluminum-containing refractory high-entropy alloys. Jom 2014, 66, 2030–2042. [Google Scholar] [CrossRef]
  17. Rizi, M.S.; Minouei, H.; Lee, B.J.; Pouraliakbar, H.; Toroghinejad, M.R.; Hong, S.I. Hierarchically activated deformation mechanisms to form ultra-fine grain microstructure in carbon containing FeMnCoCr twinning induced plasticity high entropy alloy. Mater. Sci. Eng. A 2021, 824, 141803. [Google Scholar] [CrossRef]
  18. Shim, S.H.; Pouraliakbar, H.; Hong, S.I. High strength dual fcc phase CoCuFeMnNi high-entropy alloy wires with dislocation wall boundaries stabilized by phase boundaries. Mater. Sci. Eng. A 2021, 825, 141875. [Google Scholar] [CrossRef]
  19. Senkov, O.N.; Wilks, G.B.; Miracle, D.B.; Chuang, C.P.; Liaw, P.K. Refractory high-entropy alloys. Intermetallics 2010, 18, 1758–1765. [Google Scholar] [CrossRef]
  20. Zou, Y.; Maiti, S.; Steurer, W.; Spolenak, R. Size-dependent plasticity in an nb25mo25ta25w25 refractory high-entropy alloy. Acta Mater. 2014, 65, 85–97. [Google Scholar] [CrossRef]
  21. Yao, H.; Qiao, J.-W.; Gao, M.C.; Hawk, J.A.; Ma, S.-G.; Zhou, H. MoNbTaV medium-entropy alloy. Entropy 2016, 18, 189. [Google Scholar] [CrossRef] [Green Version]
  22. Körmann, F.; Sluiter, M.H.F. Interplay between lattice distortions, vibrations and phase stability in NbMoTaW high entropy alloys. Entropy 2016, 18, 403. [Google Scholar] [CrossRef] [Green Version]
  23. Yao, H.W.; Qiao, J.W.; Hawk, J.A.; Zhou, H.F.; Chen, M.W.; Gao, M.C. Mechanical properties of refractory high-entropy alloys: Experiments and modeling. J. Alloy. Compd. 2017, 696, 1139–1150. [Google Scholar] [CrossRef]
  24. Dobbelstein, H.; Thiele, M.; Gurevich, E.L.; George, E.P.; Ostendorf, A. Direct metal deposition of refractory high entropy alloy MoNbTaW. Phys. Procedia 2016, 83, 624–633. [Google Scholar] [CrossRef] [Green Version]
  25. Kube, S.A.; Sohn, S.; Uhl, D.; Datye, A.; Mehta, A.; Schroers, J. Phase selection motifs in high entropy alloys revealed through combinatorial methods: Large atomic size difference favors BCC over FCC. Acta Mater. 2019, 166, 677–686. [Google Scholar] [CrossRef]
  26. Xu, Q.; Guan, H.Q.; Zhong, Z.H.; Huang, S.S.; Zhao, J.J. Irradiation resistance mechanism of the CoCrFeMnNi equiatomic high-entropy alloy. Sci. Rep. 2021, 11, 608. [Google Scholar] [CrossRef]
  27. McAlpine, S.W.; Logan, J.V.; Short, M.P. Predicting single phase stability and segregation in the NbMoTaTi–(W,V) high entropy alloy system with the vacancy exchange potential. Scr. Mater. 2021, 191, 29–33. [Google Scholar] [CrossRef]
  28. Chokshi, A.H. High temperature deformation in fine grained high entropy alloys. Mater. Chem. Phys. 2018, 210, 152–161. [Google Scholar] [CrossRef]
  29. Chen, W.; Ding, X.; Feng, Y.; Liu, X.; Liu, K.; Lu, Z.P.; Li, D.; Li, Y.; Liu, C.T.; Chen, X.Q. Vacancy formation enthalpies of high-entropy FeCoCrNi alloy via first-principles calculations and possible implications to its superior radiation tolerance. J. Mater. Sci. Technol. 2018, 34, 355–364. [Google Scholar] [CrossRef]
  30. Roy, A.; Munshi, J.; Balasubramanian, G. Low energy atomic traps sluggardize the diffusion in compositionally complex refractory alloys. Intermetallics 2021, 131, 107106. [Google Scholar] [CrossRef]
  31. Huang, Y.C.; Lai, Y.C.; Lin, Y.H.; Wu, S.K. A study on the severely cold-rolled and annealed quaternary equiatomic derivatives from quinary HfNbTaTiZr refractory high entropy alloy. J. Alloy. Compd. 2021, 855, 157404. [Google Scholar] [CrossRef]
  32. Cao, P.P.; Huang, H.L.; Jiang, S.H.; Liu, X.J.; Wang, H.; Wu, Y.; Lu, Z.P. Microstructural stability and aging behavior of refractory high entropy alloys at intermediate temperatures. J. Mater. Sci. Technol. 2022, 122, 243–254. [Google Scholar] [CrossRef]
  33. Antillon, E.; Woodward, C.; Rao, S.I.; Akdim, B.; Parthasarathy, T.A. Chemical short range order strengthening in a model FCC high entropy alloy. Acta Mater. 2020, 190, 29–42. [Google Scholar] [CrossRef]
  34. Ma, E. Unusual dislocation behavior in high-entropy alloys. Scr. Mater. 2020, 181, 127–133. [Google Scholar] [CrossRef]
  35. Lu, C.; Yang, T.; Jin, K.; Gao, N.; Xiu, P.; Zhang, Y.; Gao, F.; Bei, H.; William, J.; Weber, K.S.; et al. Radiation-induced segregation on defect clusters in single-phase concentrated solid-solution alloys. Acta Mater. 2017, 127, 98–107. [Google Scholar] [CrossRef]
  36. Jin, K.; Sales, B.C.; Stocks, G.M.; Samolyuk, G.D.; Daene, M.; Weber, W.J.; Bei, H. Tailoring the physical properties of Ni-based single-phase equiatomic alloys by modifying the chemical complexity. Sci. Rep. 2016, 6, 20159. [Google Scholar] [CrossRef]
  37. Sales, B.C.; Jin, K.; Bei, H.; Stocks, G.M.; Samolyuk, G.D.; May, A.F. Quantum critical behavior in a concentrated ternary solid solution. Sci. Rep. 2016, 6, 26179. [Google Scholar] [CrossRef] [Green Version]
  38. Tulip, P.R.; Staunton, J.B.; Lowitzer, S.; Ködderitzsch, D.; Ebert, H. Theory of electronic transport in random alloys with short-range order: Korringa-Kohn-Rostoker nonlocal coherent potential approximation. Phys. Rev. B 2008, 77, 165116. [Google Scholar] [CrossRef] [Green Version]
  39. Shen, Z.; Du, J.P.; Shinzato, S.; Sato, Y.; Yu, P.; Ogata, S. Kinetic Monte Carlo simulation framework for chemical short-range order formation kinetics in a multi-principal-element alloy. Comput. Mater. Sci. 2021, 198, 110670. [Google Scholar] [CrossRef]
  40. Zhao, S. Role of chemical disorder and local ordering on defect evolution in high-entropy alloys. Phys. Rev. Mater. 2021, 5, 103604. [Google Scholar] [CrossRef]
  41. Zhao, S. Effects of local elemental ordering on defect-grain boundary interactions in high-entropy alloys. J. Alloys Compd. 2021, 887, 161314. [Google Scholar] [CrossRef]
  42. Barr, C.M.; Nathaniel, J.E., II; Unocic, K.A.; Liu, J.; Zhang, Y.; Wang, Y.; Taheri, M.L. Exploring radiation induced segregation mechanisms at grain boundaries in equiatomic CoCrFeNiMn high entropy alloy under heavy ion irradiation. Scr. Mater. 2018, 156, 80–84. [Google Scholar] [CrossRef]
  43. Esfandiarpour, A.; Nasrabadi, M.N. Vacancy formation energy in CuNiCo equimolar alloy and CuNiCoFe high entropy alloy: Ab initio based study. Calphad 2019, 66, 101634. [Google Scholar] [CrossRef]
  44. Zhang, X.; Divinski, S.V.; Grabowski, B. Ab initio prediction of vacancy energetics in HCP Al-Hf-Sc-Ti-Zr high entropy alloys and the subsystems. Acta Mater. 2022, 227, 117677. [Google Scholar] [CrossRef]
  45. Li, C.; Yin, J.; Odbadrakh, K.; Sales, B.C.; Zinkle, S.J.; Stocks, G.M.; Wirth, B.D. First principle study of magnetism and vacancy energetics in a near equimolar NiFeMnCr high entropy alloy. J. Appl. Phys. 2019, 125, 155103. [Google Scholar] [CrossRef]
  46. Xing, B.; Wang, X.; Bowman, W.J.; Cao, P. Short-range order localizing diffusion in multi-principal element alloys. Scr. Mater. 2022, 210, 114450. [Google Scholar] [CrossRef]
  47. Zhao, S. Defect properties in a VTaCrW equiatomic high entropy alloy (HEA) with the body centered cubic (bcc) structure. J. Mater. Sci. Technol. 2020, 44, 133–139. [Google Scholar] [CrossRef]
  48. Luke, N. Simulating Vacancy Formation and Diffusion in NbMoTaW; University of California: San Diego, CA, USA, 2020. [Google Scholar]
  49. Thompson, A.P.; Aktulga, H.M.; Berger, R.; Bolintineanu, D.S.; Brown, W.M.; Crozier, P.S.; Kohlmeyer, A.; Moore, S.G.; Nguyen, T.D.; Shan, R.; et al. LAMMPS—A flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales. Appl. Phys. Lett. 2015, 106, 161906. [Google Scholar] [CrossRef]
  50. Li, X.-G.; Chen, C.; Zheng, H.; Zuo, Y.; Ong, S.P. Complex strengthening mechanisms in the NbMoTaW multi-principal element alloy. npj Comput. Mater. 2020, 6, 70. [Google Scholar] [CrossRef]
  51. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO–the Open Visualization Tool. Model. Simul. Mater. Sci. Eng. 2009, 18, 015012. [Google Scholar] [CrossRef]
  52. de Fontaine, D. The number of independent pair-correlation functions in multi-component systems. J. Appl. Crystallogr. 1971, 4, 15–19. [Google Scholar] [CrossRef]
  53. Fernández-Caballero, A.; Wróbel, J.S.; Mummery, P.M.; Nguyen-Manh, D. Short-range order in high entropy alloys: Theoretical formulation and application to Mo-Nb-Ta-V-W system. J. Phase Equilibria Diffus. 2017, 38, 391–403. [Google Scholar] [CrossRef] [Green Version]
  54. Byggmästar, J.; Nordlund, K.; Djurabekova, F. Modeling refractory high-entropy alloys with efficient machine-learned interatomic potentials: Defects and segregation. Phys. Rev. B 2021, 104, 104101. [Google Scholar] [CrossRef]
  55. Hossain, M.Z.; Marian, J. Stress-dependent solute energetics in W–Re alloys from first-principles calculations. Acta Mater. 2014, 80, 107–117. [Google Scholar] [CrossRef] [Green Version]
  56. Sugita, K.; Matsuoka, N.; Mizuno, M.; Araki, H. Vacancy formation enthalpy in CoCrFeMnNi high-entropy alloy. Scr. Mater. 2020, 176, 32–35. [Google Scholar] [CrossRef]
  57. Huang, E.W.; Chou, H.S.; Tu, K.N.; Hung, W.S.; Lam, T.N.; Tsai, C.W. Element effects on high-entropy alloy vacancy and heterogeneous lattice distortion subjected to quasi-equilibrium heating. Sci. Rep. 2019, 9, 14788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Eleti, R.R.; Bhattacharjee, T.; Shibata, A.; Tsuji, N. Unique deformation behavior and microstructure evolution in high temperature processing of HfNbTaTiZr refractory high entropy alloy. Acta Mater. 2019, 171, 132–145. [Google Scholar] [CrossRef]
  59. Liu, Q.; Wang, G.; Liu, Y.; Sui, X.; Chen, Y.; Luo, S. Hot deformation behaviors of an ultrafine-grained MoNbTaTiV refractory high-entropy alloy fabricated by powder metallurgy. Mater. Sci. Eng. A 2021, 809, 140922. [Google Scholar] [CrossRef]
  60. Prohaska, T.; Irrgeher, J.; Benefield, J.; Böhlke, J.K.; Chesson, L.A.; Coplen, T.B.; Ding, T.; Dunn, P.J.H.; Gröning, M.; Holden, N.E.; et al. Standard atomic weights of the elements 2021 (IUPAC Technical Report). Pure Appl. Chem. 2022, 94, 573–600. [Google Scholar] [CrossRef]
  61. Zhao, S.; Xiong, Y.; Ma, S.; Zhang, J.; Xu, B.; Kai, J.J. Defect accumulation and evolution in refractory multi-principal element alloys. Acta Mater. 2021, 219, 117233. [Google Scholar] [CrossRef]
  62. Roy, A.; Singh, P.; Balasubramanian, G.; Johnson, D.D. Vacancy formation energies and migration barriers in multi-principal element alloys. Scr. Mater. 2022, 226, 117611. [Google Scholar] [CrossRef]
  63. Wang, Z.; Liu, C.T.; Dou, P. Thermodynamics of vacancies and clusters in high-entropy alloys. Phys. Rev. Mater. 2017, 1, 043601. [Google Scholar] [CrossRef]
Figure 1. Potential energy per atom as a function of a for random equiatomic Nb-Mo-Ta-W.
Figure 1. Potential energy per atom as a function of a for random equiatomic Nb-Mo-Ta-W.
Materials 15 05468 g001
Figure 2. Energy evolution of the random Nb-Mo-Ta-W during annealing at different temperatures. The total energy of the system at the initial state is zero for easier reference.
Figure 2. Energy evolution of the random Nb-Mo-Ta-W during annealing at different temperatures. The total energy of the system at the initial state is zero for easier reference.
Materials 15 05468 g002
Figure 3. First-nearest neighbor Warren–Cowley parameter as a function of annealing temperature.
Figure 3. First-nearest neighbor Warren–Cowley parameter as a function of annealing temperature.
Materials 15 05468 g003
Figure 4. The distributions of vacancy formation energies of: (ae) cells with short range order (annealing temperature from 300 K to 1500 K) and (f) the random cell of Nb-Mo-Ta-W. The dashed line shows the position of the statistical average of each distribution.
Figure 4. The distributions of vacancy formation energies of: (ae) cells with short range order (annealing temperature from 300 K to 1500 K) and (f) the random cell of Nb-Mo-Ta-W. The dashed line shows the position of the statistical average of each distribution.
Materials 15 05468 g004
Figure 5. Effective vacancy formation energy as a function of temperature. The horizontal line represents the statistical average in the vacancy formation energy distribution of the random Nb-Mo-Ta-W bulk. The shadow region highlights the difference between effective energies and the statistical average value.
Figure 5. Effective vacancy formation energy as a function of temperature. The horizontal line represents the statistical average in the vacancy formation energy distribution of the random Nb-Mo-Ta-W bulk. The shadow region highlights the difference between effective energies and the statistical average value.
Materials 15 05468 g005
Figure 6. The minimum energy path of vacancy migration in (a) pure constituent metals and (b) the equiatomic random Nb-Mo-Ta-W. The dashed line and the shadow region indicate the initial energy and energies below, respectively.
Figure 6. The minimum energy path of vacancy migration in (a) pure constituent metals and (b) the equiatomic random Nb-Mo-Ta-W. The dashed line and the shadow region indicate the initial energy and energies below, respectively.
Materials 15 05468 g006
Figure 7. The distributions of vacancy migration energies of: (a–e) cells with short range order (annealing temperature from 300 K to 1500 K) and (f) the random cell of Nb-Mo-Ta-W. The dashed line shows the position of the statistical average of each distribution.
Figure 7. The distributions of vacancy migration energies of: (a–e) cells with short range order (annealing temperature from 300 K to 1500 K) and (f) the random cell of Nb-Mo-Ta-W. The dashed line shows the position of the statistical average of each distribution.
Materials 15 05468 g007
Figure 8. The distributions of vacancy migration energies when: (a) only one [ 111 ] direction is considered and (b) all eight directions of [ 111 ] type are considered. Dashed lines suggest positions of statistical averages.
Figure 8. The distributions of vacancy migration energies when: (a) only one [ 111 ] direction is considered and (b) all eight directions of [ 111 ] type are considered. Dashed lines suggest positions of statistical averages.
Materials 15 05468 g008
Figure 9. A cell of random Nb-Mo-Ta-W with an edge dislocation dipole. The ‘⊥’ symbols mark the exact locations of dislocation cores. The dashed circles indicate locations with varying atomic distances from dislocation cores (e.g., position ’0’ is of zero interatomic distance from the core). Colors represent different atomic species: red = Ta, blue = Nb, yellow = Mo, green = W.
Figure 9. A cell of random Nb-Mo-Ta-W with an edge dislocation dipole. The ‘⊥’ symbols mark the exact locations of dislocation cores. The dashed circles indicate locations with varying atomic distances from dislocation cores (e.g., position ’0’ is of zero interatomic distance from the core). Colors represent different atomic species: red = Ta, blue = Nb, yellow = Mo, green = W.
Materials 15 05468 g009
Figure 10. Vacancy formation energies as a function of distance from the edge dislocation core. The labeled positions in the x axis refer to the locations shown in Figure 9. The point labeled as ’random bulk’ gives the formation energy obtained in Section 3.3.
Figure 10. Vacancy formation energies as a function of distance from the edge dislocation core. The labeled positions in the x axis refer to the locations shown in Figure 9. The point labeled as ’random bulk’ gives the formation energy obtained in Section 3.3.
Materials 15 05468 g010
Figure 11. Diffusivities of Nb-Mo-Ta-W and its pure constituents as a function of temperature.
Figure 11. Diffusivities of Nb-Mo-Ta-W and its pure constituents as a function of temperature.
Materials 15 05468 g011
Table 1. Simulated equilibrium lattice parameters of the random equiatomic Nb-Mo-Ta-W alloy, the ’average’ alloy and the individual constituents metals.
Table 1. Simulated equilibrium lattice parameters of the random equiatomic Nb-Mo-Ta-W alloy, the ’average’ alloy and the individual constituents metals.
Symbol a 0 [Å]
Nb-Mo-Ta-W3.24
Nb3.33
Ta3.32
Mo3.17
W3.18
‘average’3.25
Table 2. The chemical potential energies of pure elements and the average alloy computed in the context of the pure metals. The ‘average’ represents the weighted average of chemical potential energies of Nb, Ta, Mo, and W.
Table 2. The chemical potential energies of pure elements and the average alloy computed in the context of the pure metals. The ‘average’ represents the weighted average of chemical potential energies of Nb, Ta, Mo, and W.
Symbol μ [meV]
Nb6.1
Ta13.4
Mo22.3
W21.3
‘average’15.8
Table 3. Statistical averages of vacancy formation energy distributions of Nb-Mo-Ta-W and vacancy formation energies of the pure constituents. The ‘average’ represents the weighted average of vacancy formation energies of Nb, Ta, Mo, and W.
Table 3. Statistical averages of vacancy formation energy distributions of Nb-Mo-Ta-W and vacancy formation energies of the pure constituents. The ‘average’ represents the weighted average of vacancy formation energies of Nb, Ta, Mo, and W.
Alloy Condition E V f [eV]
300 K2.54
600 K2.53
900 K2.53
1200 K2.49
1500 K2.49
random2.48
Nb2.32
Ta2.24
Mo2.81
W2.93
‘average’2.57
Table 4. The effective vacancy formation energies of Nb-Mo-Ta-W at different temperatures.
Table 4. The effective vacancy formation energies of Nb-Mo-Ta-W at different temperatures.
T [K] E eff f [eV]
15002.367
16002.375
17002.383
18002.393
20002.398
Table 5. Statistical averages of vacancy migration energy distributions of Nb-Mo-Ta-W and vacancy migration energies of the pure constituents. The ‘average’ represents the weighted average of vacancy migration energies of Nb, Ta, Mo, and W.
Table 5. Statistical averages of vacancy migration energy distributions of Nb-Mo-Ta-W and vacancy migration energies of the pure constituents. The ‘average’ represents the weighted average of vacancy migration energies of Nb, Ta, Mo, and W.
Alloy Condition E V m [eV]
300 K1.89
600 K1.84
900 K1.82
1200 K1.82
1500 K1.79
random1.71
Nb0.99
Ta1.32
Mo1.74
W2.07
‘average’1.53
Table 6. Statistical averages of vacancy formation energy distributions of position ’0’ in alloys with varying degrees of short range order. The ‘average’ represents the weighted average of vacancy formation energies at position ‘0’ in Nb, Ta, Mo, and W.
Table 6. Statistical averages of vacancy formation energy distributions of position ’0’ in alloys with varying degrees of short range order. The ‘average’ represents the weighted average of vacancy formation energies at position ‘0’ in Nb, Ta, Mo, and W.
Alloy Condition E 0 f [eV]
300 K1.11
600 K1.12
900 K1.05
1200 K1.09
1500 K1.10
random1.00
Nb0.86
Ta0.81
Mo0.94
W0.97
‘average’0.90
Table 7. Parameters and pre-exponential factors used in calculation of diffusivity.
Table 7. Parameters and pre-exponential factors used in calculation of diffusivity.
SystemK [GPa] [50]m [a.m.u.] [60] D 0 [ × 10 6 m 2 ·s 1 ]
Nb-Mo-Ta-W83138.412.96
Nb3292.912.39
Ta59180.952.31
Mo11095.953.86
W160183.843.40
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, X.; He, S.; Marian, J. Vacancy Energetics and Diffusivities in the Equiatomic Multielement Nb-Mo-Ta-W Alloy. Materials 2022, 15, 5468. https://doi.org/10.3390/ma15155468

AMA Style

Zhou X, He S, Marian J. Vacancy Energetics and Diffusivities in the Equiatomic Multielement Nb-Mo-Ta-W Alloy. Materials. 2022; 15(15):5468. https://doi.org/10.3390/ma15155468

Chicago/Turabian Style

Zhou, Xinran, Sicong He, and Jaime Marian. 2022. "Vacancy Energetics and Diffusivities in the Equiatomic Multielement Nb-Mo-Ta-W Alloy" Materials 15, no. 15: 5468. https://doi.org/10.3390/ma15155468

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