Next Article in Journal
Leaching of Copper Concentrate with Iodized Salts in a Saline Acid Medium: Part 1—Effect of Concentrations
Next Article in Special Issue
Microstructural Changes and Mechanical Properties of Precipitation-Strengthened Medium-Entropy Fe71.25(CoCrMnNi)23.75Cu3Al2 Maraging Alloy
Previous Article in Journal
Paraffin-Multilayer Graphene Composite for Thermal Management in Electronics
Previous Article in Special Issue
Phase Transformations Caused by Heat Treatment and High-Pressure Torsion in TiZrHfMoCrCo Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Estimation of Shear Modulus and Hardness of High-Entropy Alloys Made from Early Transition Metals Based on Bonding Parameters

1
H-ION Research, Development and Innovation Ltd., Konkoly-Thege Miklos út 29-33, H-1121 Budapest, Hungary
2
Department of Materials Physics, Eötvös Loránd University, Pázmány Péter Sétány 1/A, H-1117 Budapest, Hungary
3
Wigner Research Centre for Physics, Institute for Solid State Physics and Optics, Konkoly-Thege Miklos út 29-33, H-1121 Budapest, Hungary
4
Applied Materials Physics, Department of Materials Science and Engineering, Royal Institute of Technology, SE-100 44 Stockholm, Sweden
*
Author to whom correspondence should be addressed.
Materials 2023, 16(6), 2311; https://doi.org/10.3390/ma16062311
Submission received: 1 February 2023 / Revised: 27 February 2023 / Accepted: 7 March 2023 / Published: 13 March 2023
(This article belongs to the Special Issue Advances in High Entropy Materials)

Abstract

:
The relationship between the tendencies towards rigidity (measured by shear modulus, G) and hardness (measured by Vickers hardness, HV) of early transition metal (ETM)-based refractory high-entropy alloys (RHEA) and bond parameters (i.e., valence electron concentration (VEC), enthalpy of mixing (ΔHmix)) was investigated. These bond parameters, VEC and ΔHmix, are available from composition and tabulated data, respectively. Based on our own data (9 samples) and those available from the literatures (47 + 27 samples), it seems that for ETM-based RHEAs the G and HV characteristics have a close correlation with the bonding parameters. The room temperature value of G and HV increases with the VEC and with the negative value of ΔHmix. Corresponding equations were deduced for the first time through multiple linear regression analysis, in order to help design the mechanical properties of ETM refractory high-entropy alloys.

1. Introduction

The name high-entropy alloy (HEA) originated from Yeh et al. [1] almost 20 years ago. This name came from the role of the high mixing entropy, as the dominant term of the Gibbs free energy described by the equation   Δ G m i x = Δ H m i x T Δ S m i x , where Δ H m i x   and Δ S m i x   are the enthalpy and entropy of mixing, respectively, and T is the absolute temperature. In the last two decades, the study of high-entropy alloys (HEAs) has been at the forefront of materials science, due to their unique mechanical [2,3,4,5,6] and corrosion resistive [7,8,9] behaviors.
Early transition metal (ETM)-based refractory high-entropy alloys (RHEAs) containing only a single-phase body-centered cubic (BCC) structure have been intensively studied since the late 2000s [10]. In addition to their excellent mechanical properties, the RHEAs suffer from two main drawbacks: a lack of oxidation resistance when working in air at temperatures above 1000 K, and a poor ductility at room temperature in the as-cast state. In the last 10 years, many criteria for the ductile–brittle behavior of these materials have been presented [11]. It should be emphasized that all of them contain the shear modulus, G.
It is well known that shear modulus characterizes the rigidity of a sample and it is a measure of the resistance to shape change. A cube can be deformed by stretching along the diagonal of one of the faces (trigonal distortion) and along the axis (tetragonal distortion). The relevant elastic constants in these cases are called C44 and C′, respectively. The shear modulus can be expressed as a function of these elastic constants [12] as
G V = C 44 2 5 ( C 44 C )
in Voigt’s notation [12] and
G R = 5 C 44 C 2 C 44 + 3 C  
in Reuss’s notation [12], where
C = C 11 C 12 2
The G will be the arithmetic Hill average of the lower (GV) and upper (GR) limits:
G = G V + G R 2  
It should be mentioned that for an isotropic cubic alloy C44 = C′, and in this case all of elastic constants are the same, that is
GV = GR = G = C44 = C′
In general, the deviation from isotropy can be measured, for example, by using Zener’s anisotropy, AZ defined as [12]:
A Z = C 44 C
which is equal to “1” for perfect isotropy and can be as high as “10” for amorphous alloys [12].
In order to help in designing ductile RHEAs, here we propose an empirical criterion for determining the rigidity G and hardness HV by means of bond parametric functions, which comprise of VEC and ΔHmix, which are important characteristics of high-entropy alloys [13]. These parameters are defined using well-known formulas:
V E C = i n V E C i c i
where VECi is the valence electron concentration of element i with atomic concentration ci, and
Δ H = i < j 4 H i j c i c j
where Hij is the enthalpy of mixing of elements i and j at the equimolar concentration in regular binary solutions [13].

2. Materials and Methods

Metallic elements in wire and chunk form, having a purity of 99.95%, were used to produce samples of 15 g. The elongated ellipsoid shape rods about 40 mm long and with a diameter of about 10 mm were prepared by induction melting in a water-cooled copper mold under argon atmosphere. The rods were re-melted 5 times and held above the melting point to homogenize the ingot. Nine different composition samples (see Table 1) were prepared. All of these 9 samples had a single phase BCC structure, which was confirmed by XRD investigations.
Microhardness data (HV) were determined at room temperature on the mirror-polished surface perpendicular to the axis of the sample rod. The hardness measurements were carried out using a Vickers type indenter at a load of 1 kg on a Zwick/Roell-ZHμ-Indentec microhardness tester. At least ten measurements were performed on each sample, and then their average was taken as the characteristic value for the sample.
The elastic constancies C11 and C12 were determined from the bulk modulus, B, using the equation in [16]:
B = (C11+2C12)/3
and from the tetragonal component, C′, of the shear modulus described in Equation (3).
In the ab initio calculations, the bulk modulus was extracted from the Morse function fitted to the total energies, calculated as a function of volume. The total energy and the two components of the shear elastic parameters, C′ and C44, were computed according to the standard methodology of density functional theory. More details of the process can be found in Ref. [22].

3. Results and Discussion

3.1. Estimation of Shear Modulus, G, Based on Bonding Parameters

In a seminal paper [23], Saito et al. showed 20 years ago that the C′ component of G varies linearly with the VEC number, and it becomes zero (that is C11 = C12) around VEC = 4.2 for a set of Ti-X binary BCC alloys, where X may be Nb, Ta, V, or Mo element. For a Ti-based crystalline alloy (Ti-23Nb-0.7Ta-2Zr-1.2O), the VEC was tuned to the “magic” 4.2 value by alloying with oxygen.
The basic assumption of the present work was that the mentioned linearity also applies to RHEAs, which are constituted from a mixture of ETM elements. In the present work, we considered the following ETM elements, with the VEC between 3 and 6 written in parentheses after the element: Yttrium (3), Titanium (4), Zirconium (4), Hafnium (4), Vanadium (5), Niobium (5), Tantalum (5), Chromium (6), Molybdenum (6), and Tungsten (6).
In order to confirm the mentioned assumption, elastic constants of several samples containing ETM elements were determined using ab initio calculations. These values and those from the literature are listed in Table 1. Our conjecture is confirmed on Figure 1, where we have collected the elastic constant, C′, data for all the ETM-RHEA’s samples available in the literature having single-phase BCC structure, as well as the data of the nine samples prepared for the present work.
Figure 1 shows the values of C′ (in Figure 1a) and C44 (in Figure 1b) as a function of the VEC number. It can be very clearly seen that the parameter C′ visibly changed linearly with the VEC number, completely independently of the composition of the samples. This means that there is a good correlation between the C′ component of the G and the VEC number. On the contrary, a correlation cannot be observed between the C44 component and the VEC number.
Figure 2 shows the relationships between G and VEC (Figure 2a), as well as between G and ΔHmix (Figure 2b). It can be seen that none of them showed acceptable R2 values of correlation for a linear fitting. However, the analysis showed that a good multiple linear regression can be used for fitting G as a function of VEC and ΔHmix in the form:
G = ao + a1× VEC + a2× ΔHmix
where ao is constant, and a1 and a2 are proportional coefficients, which can be obtained using multiple linear regression for the data listed in Table 1.
As a result, the G for the RHEAs system can be given as Gfitted, where:
Gfitted = −110.68 + 34.87 × VEC + 0.73 × ΔHmix
According to Equation (10), the G of any ETM-based RHEA can be estimated using the VEC number and ΔHmix mixing enthalpy calculated with Equations (7) and (8), respectively.
The values of Gfitted are also listed in Table 1, together with the accepted ones. Figure 3 shows the correlation between the accepted and fitted values of G, indicating that the proposed model is rational. It can clearly be seen that a linear proportionality with a slope of 1 (function of type f(x) = x) can be fitted to the data, clearly confirming the validity of Equation (10) for an ETM-based RHEA system.
Furthermore, from Equation (10) we have:
VEC = −(0.73/34.87) × ΔHmix + (G + 110.68)/34.87
which means that at a given value of G, the VEC parameter changes linearly with the mixing enthalpy, ΔHmix. A set of straight line VEC versus ΔHmix obtained at different values of G can be seen in Figure 4.
Figure 4 illustrates the relationship between G and bonding parameters. The variable range of G caused by VEC is 20–100 GPa, with VEC increasing from 3.5 to 6.5. It seems that the effect of the average valence electron concentration is more significant than that of the enthalpy of mixing in determining the rigidity of RHEAs.

3.2. Estimation of HV, Based on Bonding Parameters

In Table 2, the HV of 36 samples, including those from the present work and the literature, is listed, together with the VEC numbers and ΔHmix calculated using Equations (7) and (8).
Using the data in Table 2, Figure 5 shows the relationships between HV and VEC (Figure 5a), as well as between HV and ΔHmix (Figure 5b).
Based on the correlations shown in Figure 5, it can be assumed that the HV can also be expressed using a linear combination of the two bonding parameters, VEC and ΔHmix. Appling the multiple linear regression, HV can be given by the following formula:
HVfitted = −122.18 + 109.75 × VEC − 11.23 × ΔHmix
According to Equation (12), the HV of any ETM-based RHEA can be estimated using the VEC number and ΔHmix mixing enthalpy calculated by Equations (7) and (8), respectively. This estimated HV is represented in Figure 6 as a function of the measured one. It can be seen that the data are gathered around a bisector, indicating the good correlation between the estimated and measured values.
It is important to note that taking the upper limiting values for the bond parameters, at VEC = 6 and ΔHmix= −10 kJ/mol, the maximal value of HV can be predicted as:
HVmax = −122.18 + 109.75 × 6−11.23 × (−10) = 649 kgf/mm2
for an ETM-based RHEA system. This HV value (13) is 6490 MPa. Considering the yield stress, σ Y as one-third value of the HV [27], that is:
σ Y H V / 3 ,
The maximum yield stress of an ETM-based RHEA system can be estimated to be about 2100–2200 MPa.
Looking at Equation (12), it is clear that the HV of RHEAs increases with an increasing VEC number and, as well as with the negative ΔHmix. Furthermore, from Equation (12) we have:
VEC = (11.23/109.75) × ΔHmix + (HV + 122.18)/109.75,
that is, at a given value of HV, the VEC parameter changes linearly with the ΔHmix. A set of straight line VEC versus ΔHmix obtained at different values of HV can be seen in Figure 7.
Figure 7 illustrates the relationship between the hardness and bonding parameters. The variable range of HV caused by VEC is 200–600 kgf/mm2 with VEC increasing from 3.6 to 6.4. It seems that, similarly to in the case of shear modulus, the effect of average valence electron concentration is more significant than that of the enthalpy of mixing in determining the hardness of RHEAs.

4. Conclusions

Our results showed that the shear modulus, G, and HV characteristics could be correlated with two easy to determine bond parameters, VEC and ΔHmix. These bond parameters can be obtained using tabulated data of the elements and composition of the alloy. The correlation of bond parameters with the accepted values of G and HV was demonstrated using multiple linear regression calculations. The accepted values of G were obtained from elastic constants C′ and C44 and those of HV were determined using experimental measurements. Considering the limits of variable range for VEC (3.75 and 6) and for ΔHmix(−10 and 8 kJ/mol), for the ETM-based HEAs, the maximal hardness that can be foreseen for a single-phase BCC structure is about 649 kgf/mm2 (6490 MPa). Taking into account the correlation between the hardness and yield stress (see Equation (14)), the maximal yield stress is expected to be around 2170 MPa. It should also be emphasized that it is possible to adapt the developed equations for all of ETM refractory high-entropy alloys, with different compositions obtainable by combination of the nine refractory elements.
It is of importance to note that the relationship between G (and HV) and the bond parameters can probably be applied to late transition metal-based HEAs as well. The corresponding proportional coefficients will be published at a later date.

Author Contributions

O.T., L.K.V. and L.V.: Conceptualization, Writing—Original draft preparation. O.T., L.K.V. and X.L.: Investigation, Data curation. O.T., L.K.V. and N.Q.C.: Formal analysis, Writing—review and editing, L.K.V., N.Q.C. and L.V.: Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Hungarian Scientific Research Fund OTKA, Grant numbers: 128229.

Data Availability Statement

The data that support the findings of this study are partly taken from the cited references and are partly the results of the authors and not available elsewhere.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. 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]
  2. Miracle, D.B.; Miller, J.D.; Senkov, O.N.; Woodward, C.; Uchic, M.D.; Tiley, J. Exploration and development of high entropy alloys for structural applications. Entropy 2014, 16, 494–525. [Google Scholar] [CrossRef]
  3. Lu, Y.; Dong, Y.; Guo, S.; Jiang, L.; Kang, H.; Wang, T.; Wen, B.; Wang, Z.; Jie, J.; Cao, Z.; et al. A Promising New Class of High-Temperature Alloys: Eutectic High-Entropy Alloys. Sci. Rep. 2014, 4, 6200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Senkov, O.N.; Scott, J.M.; Senkova, S.V.; Miracle, D.B.; Woodward, C.F. Microstructure and room temperature properties of a high-entropy TaNbHfZrTi alloy. J. Alloys Compd. 2011, 509, 6043–6048. [Google Scholar] [CrossRef]
  5. Liao, W.B. High Strength and Deformation Mechanisms of Al0.3CoCrFeNi High-Entropy Alloy Thin Films Fabricated by Magnetron Sputtering. Entropy 2019, 21, 146. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, H.; Siu, K.W.; Liao, W.; Wang, Q. In situ mechanical characterization of CoCrCuFeNi high-entropy alloy micro/nano-pillars for their size-dependent mechanical behavior. Mater. Res. Express. 2016, 3, 094002. [Google Scholar] [CrossRef]
  7. Hsu, Y.J.; Chiang, W.C.; Wu, J.K. Corrosion behavior of FeCoNiCrCux high-entropy alloys in 3.5% sodium chloride solution. Mater. Chem. Phys. 2005, 92, 112–117. [Google Scholar] [CrossRef]
  8. Huang, P.; Yeh, J.W. Multi-Principal Element Alloys with Improved Oxidation and Wear Resistance for Thermal Spray Coating. Adv. Eng. Mater. 2004, 6, 74–78. [Google Scholar] [CrossRef]
  9. Fazakas, É.; Wang, J.Q.; Zadorozhnyy, V.; Louzguine-Luzgin, D.V.; Varga, L.K. Microstructural evolution and corrosion behavior of Al25Ti 25Ga25Be25equi-molar composition alloy. Mater. Corros. 2014, 65, 691–695. [Google Scholar] [CrossRef]
  10. Senkov, O.N.; Miracle, D.B.; Chaput, K.J.; Couzinie, J.-P. Development and exploration of refractory high entropy alloys—A review. J. Mater. Res. 2018, 153, 1–37. [Google Scholar] [CrossRef] [Green Version]
  11. Thompson, R.P.; Clegg, W.J. Predicting whether a material is ductile or brittle. Curr. Opin. Solid State Mater. Sci. 2018, 22, 100–108. [Google Scholar] [CrossRef]
  12. Grimvall, G. Thermophysical Properties of Materials; Elsevier: Amsterdam, The Netherlands, 1999. [Google Scholar]
  13. Takeuchi, A.I. 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] [Green Version]
  14. Schönecker, S.; Li, X.; Wei, D.; Nozaki, S.; Kato, H.; Vitos, L.; Li, X. Harnessing elastic anisotropy to achieve low-modulus refractory high-entropy alloys for biomedical applications. Mater. Design. 2022, 215, 110430. [Google Scholar] [CrossRef]
  15. Tian, L.-Y.; Wang, G.; Harris, J.S.; Irving, D.L.; Zhao, J.; Vitos, L. Alloying effect on the elastic properties of refractory high-entropy alloys. Mater. Des. 2017, 114, 243–279. [Google Scholar] [CrossRef]
  16. Tian, F.; Varga, L.K.; Chen, N.; Shen, J.; Vitos, L. Ab initio design of elastically isotropic TiZrNbMoVx high-entropy alloys. J. All. Compd. 2014, 599, 19–25. [Google Scholar] [CrossRef]
  17. Tian, F.; Wang, D.; Shen, J.; Wang, Y. An ab initio investgation of ideal tensile and shear strength of TiVNbMo high-entropy alloy. Mater. Lett. 2016, 166, 271–275. [Google Scholar] [CrossRef]
  18. Zheng, S.; Feng, W.; Wang, S. Elastic properties of high entropy alloys by MaxEnt approach. Comput. Mater. Sci. 2018, 142, 332–337. [Google Scholar] [CrossRef]
  19. Mu, Y.K.; Liu, H.X.; Liu, Y.H.; Zhang, X.W.; Jiang, Y.H.; Dong, T. An ab initio and experimental studies of the structure, mechanical parameters and state density on the refractory high-entropy alloy systems. J. Alloys Compd. 2017, 714, 668–696. [Google Scholar] [CrossRef]
  20. Cao, P.; Ni, X.; Tian, F.; Varga, L.K.; Vitos, L. Ab initio study of AlxMoNbTiV high-entropy alloys. J. Phys. Condens. Matter. 2015, 27, 075401. [Google Scholar] [CrossRef]
  21. Fazakas, E.; Zadorozhnyy, V.; Varga, L.K.; Inoue, A.; Louzguine-Luzgin, D.V.; Tian, F.; Vitos, L. Experimental and theoretical study of Ti20Zr20Hf20Nb20X20 (X = V or Cr) refractory high-entropy alloys. Int. J. Refract. Metals Hard Mater. 2014, 47, 131–138. [Google Scholar] [CrossRef]
  22. Vitos, L. The EMTO Method and Applications in Computational Quantum Mechanics for Materials Engineers; Springer: London, UK, 2007. [Google Scholar]
  23. Takashi, S.; Tadahiko, F.; Jung-Hwan, H.; Shigeru, K.; Kazuaki, N.; Nobuaki, S.; Rong, C.; Akira, Y.; Kazuhiko, I.; Yoshiki, S.; et al. Multifunctional Alloys Obtained via a Dislocation-Free Plastic Deformation Mechanism. Science 2003, 300, 464–467. [Google Scholar]
  24. Senkov, O.N.; Senkova, S.V.; Miracle, D.B.; Woodward, C. Mechanical properties of low-density, refractory multi-principal element alloys of the Cr–Nb–Ti–V–Zr system. Mater. Sci. Eng. A 2013, 565, 51–62. [Google Scholar] [CrossRef]
  25. Wu, Y.D.; Cai, Y.H.; Chen, X.H.; Wang, T.; Si, J.J.; Wang, L.; Wang, Y.D.; Hui, X.D. Phase composition and solid solution strengthening effect in TiZrNbMoV high-entropy alloys. Mater. Des. 2015, 83, 651–660. [Google Scholar] [CrossRef]
  26. 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. All. Compd. 2017, 696, 1139–1150. [Google Scholar] [CrossRef]
  27. Tabor, D. The Hardness of Metals; Oxford University Press: London, UK, 1951. [Google Scholar]
Figure 1. The C′ component of G changes linearly (a), whereas C44 does not correlate (b) with the VEC.
Figure 1. The C′ component of G changes linearly (a), whereas C44 does not correlate (b) with the VEC.
Materials 16 02311 g001
Figure 2. G, in the function of VEC (a); and of ΔHmix (b) for the various ETM-based RHEA samples listed in Table 1.
Figure 2. G, in the function of VEC (a); and of ΔHmix (b) for the various ETM-based RHEA samples listed in Table 1.
Materials 16 02311 g002
Figure 3. The accepted and fitted values of G line up around the first bisector, the fitted value can be calculated as G = −110.68 + 34.87 × VEC + 0.73 × ΔHmix with a goodness of fit R2 value of 0.83.
Figure 3. The accepted and fitted values of G line up around the first bisector, the fitted value can be calculated as G = −110.68 + 34.87 × VEC + 0.73 × ΔHmix with a goodness of fit R2 value of 0.83.
Materials 16 02311 g003
Figure 4. Relationship between G and bond parameters for ETM–RHEAs, indicated by linear VECHmix functions obtained at different G values.
Figure 4. Relationship between G and bond parameters for ETM–RHEAs, indicated by linear VECHmix functions obtained at different G values.
Materials 16 02311 g004
Figure 5. HV, in the function of VEC (a); and of ΔHmix (b) for the various ETM-based RHEA samples listed in Table 2.
Figure 5. HV, in the function of VEC (a); and of ΔHmix (b) for the various ETM-based RHEA samples listed in Table 2.
Materials 16 02311 g005
Figure 6. Relationship between the accepted and fitted values of HV, which line up around the first bisector.
Figure 6. Relationship between the accepted and fitted values of HV, which line up around the first bisector.
Materials 16 02311 g006
Figure 7. Relationship between HV and bond parameters for ETM–RHEAs, indicated by linear VEC- ΔHmix functions obtained at different HV values.
Figure 7. Relationship between HV and bond parameters for ETM–RHEAs, indicated by linear VEC- ΔHmix functions obtained at different HV values.
Materials 16 02311 g007
Table 1. Ab initio calculated elastic constants (in GPa), valence electron concentration, enthalpy of mixing (in kJ/mole), and shear modulus (in GPa) for the samples of the present work (pw) and those taken from the literature.
Table 1. Ab initio calculated elastic constants (in GPa), valence electron concentration, enthalpy of mixing (in kJ/mole), and shear modulus (in GPa) for the samples of the present work (pw) and those taken from the literature.
AlloysRefVECC11C12C44C′ΔHmixGGfitted
Y25Ti25Zr25Hf25pw3.7570.4085.9467.48−7.778.756.9526.21
Ti33.33Zr33.33Hf33.34pw4.0099.9994.6774.762.660.0026.1228.80
Ti45Zr45Nb5Ta5[14]4.10112.0087.0065.0012.500.9034.1332.92
Ti30Zr30Hf30Nb10pw4.10114.8894.3470.8310.272.6433.8534.14
Ti26.3Zr26.3Hf26.3Nb10.55Ta10.55[14]4.21145.00100.0073.0022.502.6845.6338.00
Ti25Zr25Hf25Nb12.5Ta12.5[14]4.25151.00101.0072.0025.002.6947.1539.40
Ti37.5Zr25Ta12.5Hf12.5Nb12.5[14]4.25146.0099.0068.0023.501.8844.4538.83
Ti25Zr25Hf25Nb25pw4.25136.2095.0664.4820.572.5040.8539.27
Ti23.8Zr23.8Hf23.8Nb14.3Ta14.3[14]4.29156.00102.0071.0027.002.3148.1940.53
Ti35Zr35Nb25Ta5[14]4.30142.0091.0058.0025.502.3841.7140.93
Ti21.7Zr21.7Hf21.7Nb17.45Ta17.45[14]4.35164.00104.0070.0030.002.5749.8342.80
Ti20Zr20Hf20Nb20V20pw4.40149.7094.2058.3527.750.1643.3042.86
NbTiVZr[15]4.50166.1093.8052.2036.15−0.2545.0546.06
NbTiVZr[16]4.50166.4094.7053.8035.85−0.2545.7246.06
Ti25Nb25Ta25Zr25[14]4.50174.00101.0060.0036.502.5049.1647.99
Ti25Zr25V25Nb25pw4.50165.2692.2249.8036.52−0.2543.9846.06
Ti25Nb25Ta25Zr25[14]4.50174.00101.0060.0036.502.5049.1647.99
Ti30.5Zr30.5Nb13Ta13Mo13[14]4.52175.0097.0057.0039.00−0.7448.9646.41
Ti34Zr20Nb20Ta20Mo6[14]4.52180.00102.0060.0039.000.7950.4847.49
Ti21.67Zr21.67Nb21.66Ta35[14]4.57187.00107.0063.0040.002.3452.5150.31
Ti23.8Zr23.8Hf23.8Cr4.8Mo23.8[14]4.57192.00106.0061.0043.00−4.4553.0345.56
Ti30Zr20Nb20Ta20Mo10[14]4.60192.00104.0058.0044.000.0251.9349.74
Ti20Zr20V20Nb20Ta20pw4.60185.98102.3656.0841.810.3249.8649.95
Ti39.4Nb15.15Ta15.15Zr15.15Mo15.15[14]4.61195.00104.0058.0045.50−1.1652.6349.26
Mo0.8NbTiZr[16]4.68199.0098.7052.8050.15-1.8851.7251.20
TiZrNbMo0.8[17]4.68199.0098.7052.8050.15−1.8851.7251.20
Mo0.8NbTiV0.2Zr[16]4.70200.8099.0052.5050.90−2.0551.8551.77
Ti15Zr15Nb35Ta35[14]4.70210.00112.0060.0049.002.1055.3354.68
TiZrNbMo0.8V0.2[17]4.70200.8099.0052.5050.90−2.0551.8551.77
Mo0.9NbTiZr[16]4.72204.3099.5052.6052.40−2.2152.5252.36
TiZrNbMo0.9[17]4.72204.3099.5052.6052.40−2.2152.5252.36
Mo0.8NbTiV0.5Zr[16]4.72203.70100.0051.9051.85−2.2351.8852.35
TiZrNbMo0.8V0.5[17]4.72203.70100.0051.9051.85−2.2351.8852.35
MoNbTiZr[15]4.75209.8099.9051.3054.95−2.5052.7353.20
MoNbTiZr[16]4.75209.90101.0052.6054.45−2.5053.3353.20
TiZrNbMo[17]4.75209.90101.0052.6054.45−2.5053.3353.20
MoNbTiV0.25Zr[16]4.76211.00100.6052.1055.20−2.6053.3253.48
TiZrNbMoV0.25[17]4.76211.00100.6052.1055.20−2.6053.3253.48
MoNbTiV0.5Zr[16]4.78212.20100.3051.6055.95−2.6753.3054.13
TiZrNbMoV0.50[17]4.78212.20100.3051.6055.95-2.6753.3054.13
MoNbTiV0.75Zr[16]4.79213.20100.3051.2056.45−2.7053.2454.46
TiZrNbMoV0.75[17]4.79213.20100.3051.2056.45−2.7053.2454.46
MoNbTiVZr[16]4.80213.70100.7050.9056.50−2.7253.0754.79
MoNbTiVZr[15]4.80215.00100.5049.4057.25−2.7252.4054.79
MoNbTiVZr[18]4.80231.00105.6036.9062.70−2.7245.7054.79
TiZrNbMoV1.00[17]4.80213.70100.7050.9056.50−2.7253.0754.79
MoNbTiV1.25Zr[16]4.81218.00101.9050.0058.05−2.7253.0855.14
TiZrNbMoV1.25[17]4.81218.00101.9050.0058.05−2.7253.0855.14
MoNbTiV1.5Zr[16]4.82219.30102.2049.8058.55−2.7153.1355.50
TiZrNbMoV1.50[17]4.82219.30102.2049.8058.55−2.7153.1355.50
CrMoNbTaTiVZr[19]5.00261.40115.1038.4073.15−4.4149.8560.58
MoNbTiV[20]5.00265.50114.0051.7075.75−2.7560.2761.75
Ti25V25Nb25Mo25pw5.00264.13112.1145.4576.01−2.7555.9261.75
CrMoNbTaTiVWZr[19]5.13286.10121.9054.7082.10−5.2564.3964.53
MoNbTaVW[18]5.40392.70160.1057.60116.30−1.8176.6376.35
MoNbTaVW[21]5.40380.80177.3061.20101.75−1.8175.1176.35
V25Nb25Mo25W25pw5.50390.15131.9767.35129.09−4.0087.6678.31
Table 2. HV data (kgf/mm2) together with the VEC and ΔHmix (kJ/mole) for the samples of the present work and those taken from the literature.
Table 2. HV data (kgf/mm2) together with the VEC and ΔHmix (kJ/mole) for the samples of the present work and those taken from the literature.
AlloysRefVECHVΔHmixHV Fitted
Nb28.3Ti24.5V23Zr24.2[24]4.513350.05376
Nb22.6Ti19.4V37.2Zr20.8[24]4.60352−1.05397
Cr24.6Nb26.7Ti23.9Zr24.8[24]4.76418−4.84451
Cr20.2Nb20Ti19.9V19.6Zr20.3[24]4.80481−4.68454
Y25Ti25Zr25Hf25pw3.752158.75204
Ti33.33Zr33.33Hf33.34pw4.002980.00315
Ti30Zr30Hf30Nb10pw4.103331.20316
Ti25Zr25Hf25Nb25pw4.253362.50322
Ti20Zr20Hf20Nb20V20pw4.403920.16362
Ti25Zr25V25Nb25pw4.50385−0.25377
Ti20Zr20V20Nb20Ta20pw4.604100.32384
Ti25V25Nb25Mo25pw5.00470−2.75461
V25Nb25Mo25W25pw5.50472−4.00532
TiZrNbV[25]4.50325−0.25377
TiZrNbVMo0.3[25]4.61379−1.26399
TiZrNbVMo0.5[25]4.67433−1.78412
TiZrNbVMo0.7[25]4.72450−2.21422
TiZrNbVMo1.0[25]4.80460−2.72436
TiZrNbVMo1.3.[25]4.87440−3.10448
TiZrNbVMo1.5[25]4.91472−3.31455
TiZrNbVMo1.7[25]4.95484−3.47461
TiZrNbVMo2.0[25]5.00519−3.67469
TiZrNbV0.3[25]4.393041.43349
TiZrNbV0.3Mo0.1[25]4.443300.80361
TiZrNbV0.3Mo0.3[25]4.53386−0.28381
TiZrNbV0.3Mo0.5[25]4.61383−1.14398
TiZrNbV0.3Mo0.7[25]4.68422−1.83413
TiZrNbV0.Mo1.03[25]4.78428−2.62432
TiZrNbV0.3Mo1.3[25]4.85472−3.20447
TiZrNbV0.3Mo1.5[25]4.90464−3.49455
NbCrMo0.5Ta0,5TiZr[25]4.90469−4.92469
MoNbTaV[26]5.25504−3.25495
NbTaVW[26]5.25493−4.50507
NbTaTiVW[26]5.00447−3.68469
MoNbTaVW[26]5.40535−4.64526
NbTiVZr[26]4.50335−0.25377
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

Temesi, O.; Varga, L.K.; Li, X.; Vitos, L.; Chinh, N.Q. Estimation of Shear Modulus and Hardness of High-Entropy Alloys Made from Early Transition Metals Based on Bonding Parameters. Materials 2023, 16, 2311. https://doi.org/10.3390/ma16062311

AMA Style

Temesi O, Varga LK, Li X, Vitos L, Chinh NQ. Estimation of Shear Modulus and Hardness of High-Entropy Alloys Made from Early Transition Metals Based on Bonding Parameters. Materials. 2023; 16(6):2311. https://doi.org/10.3390/ma16062311

Chicago/Turabian Style

Temesi, Ottó, Lajos K. Varga, Xiaoqing Li, Levente Vitos, and Nguyen Q. Chinh. 2023. "Estimation of Shear Modulus and Hardness of High-Entropy Alloys Made from Early Transition Metals Based on Bonding Parameters" Materials 16, no. 6: 2311. https://doi.org/10.3390/ma16062311

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